Skip to main content

Full text of "NC 8th Grade Chemistry"

See other formats


CK-12 Foundation 



NC 8th Grade Chemistry 



Say Thanks to the Authors 

Click http://www.ckl2.org/saythanks 
(No sign in required) 



Bewick Edge Forsythe Parsons 



To access a customizable version of this book, as well as other interactive content, visit www.ckl2.org 



CK-12 Foundation is a non-profit organization with a mission to reduce the cost of textbook mate- 
rials for the K-12 market both in the U.S. and worldwide. Using an open-content, web-based collaborative 
model termed the FlexBook®, CK-12 intends to pioneer the generation and distribution of high-quality 
educational content that will serve both as core text as well as provide an adaptive environment for 
learning, powered through the FlexBook Platform®. 

Copyright © 2011 CK-12 Foundation, www.ckl2.org 

The names "CK-12" and "CK12" and associated logos and the terms "FlexBook®", and "FlexBook 
Platform®", (collectively "CK-12 Marks") are trademarks and service marks of CK-12 Foundation and 
are protected by federal, state and international laws. 

Any form of reproduction of this book in any format or medium, in whole or in sections must include the 
referral attribution link http://www.ckl2.org/saythanks (placed in a visible location) in addition to 
the following terms. 

Except as otherwise noted, all CK-12 Content (including CK-12 Curriculum Material) is made available 
to Users in accordance with the Creative Commons Attribution/Non-Commercial/Share Alike 3.0 Un- 
ported (CC-by-NC-SA) License (http://creativecommons.0rg/licenses/by-nc-sa/3.O/), as amended 
and updated by Creative Commons from time to time (the "CC License"), which is incorporated herein 
by this reference. 

Complete terms can be found at http://www.ckl2.org/terms. 

Printed: September 11, 2011 



/lexboo< 

next generation textbooks 




Authors 

Sharon Bewick, Jonathan Edge, Therese Forsythe, Richard Parsons 



www.ckl2.org 



Contents 



1 The Science of Chemistry 1 

1.1 The Scientific Method 1 

1.2 Chemistry in History 29 

1.3 Chemistry is a Science of Materials 35 

1.4 Matter 42 

1.5 Energy 48 

2 Chemistry - A Physical Science 57 

2.1 Measurements in Chemistry 57 

2.2 Using Measurements 62 

2.3 Using Mathematics in Chemistry 66 

2.4 Using Algebra in Chemistry 69 

3 Chemistry in the Laboratory 75 

3.1 Making Observations 75 

3.2 Making Measurements 78 

3.3 Using Data 89 

3.4 How Scientists Use Data 110 

4 Atomic Theory 115 

4.1 Early Development of a Theory 115 

4.2 Further Understanding of the Atom 123 

4.3 Atomic Terminology 133 

5 The Bohr Model 148 

5.1 The Wave Form of Light 148 

5.2 The Dual Nature of Light 154 

5.3 Light and the Atomic Spectra 164 

5.4 The Bohr Model 171 

www.ckl2.org 11 



6 Quantum Mechanics Model of 

the Atom 182 

6.1 The Wave Particle Duality 182 

6.2 Schrodinger's Wave Functions 191 

6.3 Heisenberg's Contribution 198 

6.4 Quantum Numbers 203 

6.5 Shapes of Atomic Orbitals 213 

7 Electron Configurations for 

Atoms 224 

7.1 The Electron Spin Quantum Number 224 

7.2 Pauli Exclusion Principle 233 

7.3 Aufbau Principle 236 

7.4 Writing Electron Configurations 240 

8 Electron Configurations and 

the Periodic Table 246 

8.1 Electron Configurations of Main Group Elements 246 

8.2 Orbital Configurations 257 

8.3 The Periodic Table and Electron Configurations 263 

9 Relationships Between the 

Elements 272 

9.1 Families on Periodic Table 272 

9.2 Electron Configurations 281 

9.3 Lewis Electron Dot Diagrams 291 

9.4 Chemical Family Members Have Similar Properties 297 

9.5 Transition Elements 300 

9.6 Lanthanides and Actinides 305 

10 Trends on the Periodic Table 310 

10.1 Atomic Size 310 

10.2 Ionization Energy 321 

10.3 Electron Affinity 327 

11 Covalent Bonding 331 

11.1 The Covalent Bond 331 

11.2 Atoms that Form Covalent Bonds 335 

11.3 Naming Covalent Compounds 342 

111 www.ckl2.org 



12 Reactions 345 

12.1 Chemical Equations 345 

12.2 Balancing Equations 349 

12.3 Types of Reactions 356 

13 The Liquid State 372 

13.1 The Properties of Liquids 372 

13.2 Forces of Attraction 374 

13.3 Vapor Pressure 382 

13.4 Boiling Point 389 

13.5 Heat of Vaporization 392 

14 The Solid State 396 

14.1 The Molecular Arrangement in Solids Controls Solid Characteristics 396 

14.2 Melting 397 

14.3 Types of Forces of Attraction for Solids 402 

14.4 Phase Diagrams 410 

15 The Solution Process 415 

15.1 What Are Solutions? 415 

15.2 Why Solutions Occur 418 

15.3 Solution Terminology 421 

15.4 Measuring Concentration 424 

15.5 Solubility Graphs 429 

15.6 Factors Affecting Solubility 435 

15.7 Colligative Properties 439 

15.8 Colloids 445 

15.9 Separating Mixtures 447 

16 Ions in Solution 452 

16.1 Ionic Solutions 452 

16.2 Covalent Compounds in Solution 458 

16.3 Reactions Between Ions in Solutions 462 

17 Acids and Bases 481 

17.1 Arrhenius Acids 481 

17.2 Strong and Weak Acids 486 

17.3 Arrhenius Bases 489 

17.4 Salts 493 

www.ckl2.org IV 



17.5 pH 502 

17.6 Weak Acid/Base Equilibria 510 

17.7 Br0nsted Lowry Acids-Bases 516 

17.8 Lewis Acids and Bases 522 

18 Water, pH and Titration 527 

18.1 Water Ionizes 527 

18.2 Indicators 531 

18.3 Titrations 541 

18.4 Buffers 559 

19 Radioactivity and the Nucleus 564 

19.1 Discovery of Radioactivity 564 

19.2 Nuclear Notation 567 

19.3 Nuclear Force 569 

19.4 Nuclear Disintegration 572 

19.5 Nuclear Equations 576 

19.6 Radiation Around Us 581 

19.7 Applications of Nuclear Energy 585 



www.ckl2.org 



www.ckl2.org VI 



Chapter 1 

The Science of Chemistry 



1.1 The Scientific Method 

Lesson Objectives 

• Describe the steps involved in the scientific method. 

• Appreciate the value of the scientific method. 

• Recognize that in some cases not all the steps in the scientific method occur, or they do not occur in 
any specific order. 

• Explain the necessity for experimental controls. 

• Recognize the components in an experiment that represent experimental controls. 

Introduction 

"What hopes and fears does this scientific method imply for mankind? I do not think that this is the right 
way to put the question. Whatever this tool in the hand of man will produce depends entirely on the 
nature of the goals alive in this mankind. Once the goals exist, scientific method furnishes means to realize 
them. Yet it cannot furnish the very goals. The scientific method itself would not have led anywhere, it 
would not even have been born without a passionate striving for clear understanding." - Albert Einstein 



Historical Comparisons 
Introduction to Science 

What is science? Is it a list of marvelous inventions and how they work? Or is it a list of theories about 
matter and energy and biological systems? Or is science a subject that you learn by carrying out activities 
in a laboratory? Science is all of these, but it is also something even more basic. Science is a method of 
thinking that allows us to discover how the world around us works. 

To begin this study of one form of science, we will review the last 3,000 years in the history of human 
transportation, communication, and medicine. The following summary lists humankind's accomplishments 
in these areas during three periods in the last 3,000 years. 

1 www.ckl2.org 



Transportation in 1000 B.C. 

In 1000 B.C., people could transport themselves and their goods by walking, riding an animal, or by riding 
in a cart pulled by an animal (Figure ??). Crossing water, people could paddle a boat or have an animal 
walk beside the river and pull the boat (Figure 1.1). These methods of transportation required muscle 
power, either human muscles or animal muscles. 





Figure 1.1: A photo of a wooden model of a Greek ship that has both sails and oars. 

A few societies had designed rowboats or sailboats, which used muscle power or the force of the wind to 
move the boat. These early means of transportation were very limited in terms of speed and therefore, 
also limited in terms of distances traveled. The sail and rowboats were used on rivers and inland seas, but 
were not ocean-going vessels. 

Transportation in 1830 

By the year 1830, people were still walking and riding in carts pulled by animals. Iron ore was moved 
along canals by animals pulling barges. American pioneers crossed the United States in covered wagons 
pulled by animals (Figure 1.2). Large cities had streetcars pulled by horses (Figure 1.3). Ocean crossing 
was accomplished in sailing ships. The only improvement in transportation was the addition of springs 



www.ckl2.org 



and padded seats to carts and wagons to make the ride less jolting. In the period from 1000 B.C. to 1830, 
a span of 2,830 years (about 100 generations of people), there were no significant changes in the mode 
of human transportation. 




Figure 1.2: A covered wagon of the type used by pioneers to cross the US in the mid- 1800s. 




Figure 1.3: The first horse-drawn street car in Seattle, Washington in 1884. 



Transportation in 1995 

By the year 1995, steam engines, gasoline engines, automobiles, propeller-driven and jet engines, loco- 
motives, nuclear-powered ships, and inter-planetary rocket ships were invented (Figure 1.4). In all in- 
dustrialized countries, almost anyone could own an automobile and travel great distances in very short 
times. 

In the mid- 1800s, several months were required to travel from Missouri to California by covered wagon 
and the trip was made at considerable risk to the traveler's life. In 1995, an average family could travel 
this same distance easily in two days and in relative safety. An ordinary person in 1995 probably traveled 
a greater distance in one year than an ordinary person in 1830 did in an entire lifetime. The significant 
changes in the means of transportation in the 165 years between 1830 and 1995 (perhaps 5 generations) 
were phenomenal. 



www.ckl2.org 




Figure 1.4: A modern jetliner. 



Communication in 1000 B.C. 




Figure 1.5: Shopping list chiseled on a rock. 

Essentially, people's only means of communicating over large distances (more than 15 miles) in 1000 B.C. 
was to send hand-carried messages (Figure 1.5). Some societies, for short distances, had developed the use 
of smoke signals, light signals, or drum signals, but these methods were useless for long distances. Since 
the means of communicating required hand-carried messages, the speed of communication was limited by 
the speed of transportation. Sending messages over distances of 1,000 miles could require several weeks 
and even then delivery was not guaranteed. 



Communication in 1830 

By the year 1830, people's means of communication over large distances was still the hand-carried message. 
While the paper and ink used to write the message had been improved, it still had to be hand-carried. 
In the United States, communication between New York and San Francisco required more than a month. 
When a new president was elected, Californians would not know who it was for a couple of months after 
the election. 

For a short period of time, the Pony Express was set up and could deliver a letter from St. Louis, Missouri 
to Sacramento, California in eleven days, which was amazing at the time (Figure 1.6). The means of 
communication in 1830 was essentially the same as in 1000 B.C. 



www.ckl2.org 




Figure 1.6: A pony express rider, circa 1861. 



Communication in 1995 




Figure 1.7: A modern cell phone. 

By the year 1995, the telegraph, telephone, radio, television, optical fibers, and communication satellites 
were invented (Figure 1.7). People could communicate almost anywhere in the industrialized world in- 
stantaneously. Now, when a U.S. president is elected, people around the globe know the name of the new 
president the instant the last vote is counted. Astronauts communicate directly between the earth and 
the moon. An ordinary person in an industrialized country can speak with people around the world while 
simultaneously watching events occur in real time globally. There have been truly extraordinary changes 
in people's ability to communicate in the last 165 years. 



www.ckl2.org 



Medical Treatment in 1000 B.C. 




Figure 1.8: Physician letting blood from a patient. 

Medical treatment in 1000 B.C. consisted of a few natural herbs and some superstitious chants and dances. 
The most advanced societies used both sorcerers and herbalists for medical treatment. Some of the natural 
herbs helped the patient and some did not. Cleaning and bandaging wounds decreased opportunity for 
infection while some herbs such as sesame oil demonstrated moderate antiseptic properties. Dances, chants, 
incense burning, and magic spells were absolutely useless in curing illnesses. At some point in time, 
bloodletting was added to the physician's repertoire (Figure 1.8). Bloodletting was accomplished by 
cutting the patient and allowing the blood to drip out or by applying leeches (which doctors often carried 
with them). However, bloodletting was not helpful to the patient, and in many cases, it was harmful. 
Bloodletting was flourishing by 500 B.C. and was carried out by both surgeons and barbers. It wasn't 
until around 1875 that bloodletting was established as quackery. 

In those times, for an ordinary person, broken bones went unset and injuries like deep cuts or stab wounds 
were often fatal due to infection. Infant mortality was high and it was common for at least one child in a 
family die before adulthood. The death of the mother in childbirth was also quite common. 

In the Middle Ages, knowledge of germs, hygiene, and contagion was non-existent. People who were 
seriously ill might have their disease blamed on the planets going out of line (astrology) or "bad odors," 
or retribution for sins, or an imbalance in body fluids. Cures could involve anything from magic spells, 
bleeding, sweating, and vomiting to re-balance bodily fluids. Between 1340 AD and 1348 AD, the Black 
Death (bubonic plague) was responsible for killing in the vicinity of half the population of Europe. The 
bacterium causing the disease was carried by fleas, but, of course, none of this was known by the physicians 
of the time. Efforts to stop the plague included burning incense to eliminate "bad odors," causing loud 
noises to chase the plague away (the constant ringing of bells or firing of canon), and a number of people 
used self-flagellation to attempt to cure the disease. 



www.ckl2.org 



6 



Medical Treatment in 1830 

Medical treatment in 1830 remained in the form of natural herbs and bloodletting. During this time, the 
ability to set broken bones and to amputate limbs was also developed. Amputation saved many lives from 
infection and gangrene. Gangrene occurs when the blood supply to tissue is interrupted and the tissue 
dies. The dead tissue remains part of the body, invites infection, and causes death as the poisons from the 
rotting tissue are carried through the body. Once gangrene afflicted an arm or leg, the poison from the 
limb would eventually kill the patient. During the American Civil War (1861 - 1863), a common means 
of treatment for wounds in field hospitals was amputation. Along with amputation was the ability to 
cauterize wounds to stop bleeding. 

Even though bloodletting did not help patients, it continued in use through 1830. There is a tale (which 
may or may not be true) that George Washington was suffering from pneumonia and his doctors removed 
so much blood trying to cure him that they actually caused his death. 

Medical Treatment in 1995 




Figure 1.9: Receiving a vaccination. 

By 1995, medical science had discovered chemical medicines, antiseptic procedures, surgery, and probably 
most important of all, vaccination . . . the ability to prevent disease rather than cure it after it had been 
contracted (Figure 1.9). 

Diseases that had killed and crippled hundreds of thousands of people in the past are seldom heard of 
today (polio, smallpox, cholera, bubonic plague, etc.). These diseases have been controlled by scientific 
understanding of their causes and carriers and by vaccination. Average life expectancy has nearly doubled 
in the last 165 years. Both infant mortality and death during childbirth rates have dropped to less than 
25% of what they were in 1830. 

Methods of Learning About Nature 
Opinion, Authority, and Superstition 

Why did humans make so little progress in the 2,800 years before 1830 and then such incredible progress 
in the 160 years after 1830? 

7 www.ckl2.org 



Socrates (469 B.C. - 399 B.C.), Plato (427 B.C. - 347 B.C.), and Aristotle (384 B.C. - 322 B.C.) are among 
the most famous of the Greek philosophers. Plato was a student of Socrates and Aristotle was a student of 
Plato. These three were probably the greatest thinkers of their time. Aristotle's views on physical science 
profoundly shaped medieval scholarship and his influence extended into the Renaissance (14'' 1 century). 
Aristotle's opinions were the authority on nature until well into the 1300s. 

Unfortunately, many of Aristotle's opinions were wrong. It is not intended here to denigrate Aristotle's 
intelligence; he was without doubt a brilliant man. It was simply that he was using a method for determining 
the nature of the physical world that is inadequate for that task. The philosopher's method was logical 
thinking, not making observations on the natural world. This led to many errors in Aristotle's thinking on 
nature. Let's consider just two of Aristotle's opinions as examples. 

In Aristotle's opinion, men were bigger and stronger than women, and therefore, it was logical that men 
would have more teeth than women. Therefore, Aristotle concluded it was a true fact that men had more 
teeth than women. Apparently, it never entered his mind to actually look into the mouths of both genders 
and count their teeth. Had he done so, he would have found that men and women have exactly the same 
number of teeth. 



QS> 




Figure Galileo dropping balls off the Leaning Tower of Pisa. 

In terms of physical science, Aristotle thought about dropping two balls of exactly the same size and shape 
but of different masses to see which one would strike the ground first. In his mind, it was clear that the 
heavier ball would fall faster than the lighter one and he concluded that this was a law of nature. Once 
again, he did not consider doing an experiment to see which ball fell faster. It was logical to him, and 
in fact, it still seems logical. If someone told you that the heavier ball would fall faster, you would have 
no reason to disbelieve it. In fact, it is not true and the best way to prove this is to try it. Eighteen 
centuries later, Galileo decided to actually get two balls of different masses, but with the same size and 
shape, and drop them off a building (legend says the Leaning Tower of Pisa), and actually see which one 
hit the ground first. When Galileo actually did the experiment, he discovered, by observation, that the 
two balls hit the ground at exactly the same time . . . Aristotle's opinion was, once again, wrong. 

In the 16 and 17 r/l centuries, innovative thinkers were developing a new way to discover the nature of the 
world around them. They were developing a method that relied upon making observations of phenomena 
and insisting that their explanations of the nature of the phenomena corresponded to the observations they 
made. In order to do this, they had to overcome the opinions of the ancient Greeks, the authority of the 
church, and the superstitions of ordinary people. 

www.ckl2.org 8 



In the opinion of the ancient Greeks, the earth was the center of the universe and did not move, while the 
sun, moon, planets, and stars revolved around the Earth in orbits. The astronomer Ptolemy (around 150 
A.D.) observed the positions of the planets and recognized that the positions where he observed the planets 
did not match up with the positions predicted by the orbits of the Greeks. Ptolemy designed new orbits 
that had circles within circles and complicated retrograde motion (planets moving backward in their orbits 
at certain times). His descriptions came closer but still could not accurately predict where the heavenly 
orbs would be on a given night. It wasn't until Nicolaus Copernicus (1473 - 1543) suggested a heliocentric 
(sun-centered) system that the positions of the planets came close to matching predictions. Copernicus was 
hesitant to publish his ideas - some say because he feared ridicule from his peers and others say because 
he feared persecution by the church - but eventually, he sent his work for publication just before his death. 

The publication of Copernicus' heliocentric theory didn't seem to cause much controversy for the next 50 
years until the idea was supported by a scientist named Giordano Bruno who was promptly prosecuted 
and burned at the stake by Cardinal Bellarmini in 1600. The most famous supporter of the Copernican 
system was Galileo Galilei (1564 - 1642) who had developed an improved telescope (1610) and turned it 
toward the sky. Galileo published a small work describing what he saw with his telescope and how his 
observations supported the Copernican theory. The book was banned by the church in 1616 and Galileo 
was instructed not to write about the subject any further. In 1632, Galileo published another work, again 
supporting the Copernican theory and was arrested by the church, prosecuted, and punished by house 
arrest for the remainder of his life. 

But the method of learning by experimenting, observing, and hypothesizing had been launched and many 
scientists would not turn back. It should be mentioned that the supporters of the methods of opinion, 
authority, and superstition did not give and have not given up today. We still have "scientists" claiming 
that unsupported opinions are "facts" and we still have people deciding the "truth" about nature by 
voting on it. Nor has superstition died. You may remember from your history classes that the pilgrims 
of Massachusetts were still drowning and hanging women accused of being witches as late as 1693. It is 
easy to think that the people of those times were not very smart, and nowadays, no one would think that 
way. However, you should be aware that a student was suspended from school in Tulsa, Oklahoma in 1999 
for "casting spells" and a substitute teacher in Florida was fired in 2008 for "wizardry" after performing a 
magic trick for his students. 

The Scientific Method 

Scientists frequently list the scientific method as a series of steps. Other scientists oppose this listing of 
steps because not all steps occur in every case and sometimes the steps are out of order. The scientific 
method is listed in a series of steps in Table 1.1 and represented in the flowchart (Figure ??) because it 
makes it easier to study. You should remember that not all steps occur in every case nor do they always 
occur in order. 

Table 1.1: The Steps in the Scientific Method 

Step Number Step Description 

1 Identify the problem or phenomenon that needs ex- 
plaining. This is sometimes referred to as "defining 
the problem." This activity helps limit the field of 
observations. 

2 Gather and organize data on the problem. This 

step is also known as "making observations." 

3 Suggest a possible solution or explanation. A sug- 
gested solution is called a hypothesis. 

9 www.ckl2.org 



Table 1.1: (continued) 



Step Number 



Step Description 



4 
5 



Test the hypothesis by making new observations. 
If the new observations support the hypothesis, you 
accept the hypothesis for further testing. If the 
new observations do not agree with your hypothe- 
sis, you discard the hypothesis, add the new data 
to your observations list, and return to step 3. 



Identify the Problem 



Gather Data 



Make a Hypothesis 



Test the Hypothesis (Experiment) 



NO 



Does the New Data Agree? 




When the results of several experiments support the hypothesis, you might think that the work is finished. 
However, for a hypothesis to be useful, it must withstand repeated testing. Other scientists must be able 
to repeat the experiments using the same materials and conditions and get the same results. Scientists 
submit reports of research to other scientists, usually by publishing an article in a scientific journal, so the 
work can be verified. 



www.ckl2.org 



10 



An Example of the Scientific Method 




Suppose you are required to maintain a large campfire and you are completely unfamiliar with the property 
of objects that makes them combustible (able to burn). The first step in the scientific method is to define 
the problem. What property of objects make them combustible? 

The next step is to gather data on the problem. So, you begin to collect objects at random and put them 
into the fire. You must keep good records of what objects were tried and whether or not they burned. 
Table 1.2 shows a list of organized data (observations): 

Table 1.2: 

Will Burn Won't Burn 

tree limbs rocks 

chair legs bricks 

pencils marbles 

baseball bat hubcaps 

The list of organized observations helps because now you can collect only the items on the "will burn" list 
and not waste the effort of dragging items that won't burn back to the fire. However, you would soon use 
up all the items on the "will burn" list and it is necessary to guess what property the "will burn" objects 
have that cause them to burn. If you had that answer, you could bring objects that may not be on the 
"will burn" list but that have the "will burn" property and keep the fire going. 

The third step in the scientific method is to suggest a hypothesis. Your guess about what property the "will 
burn" objects have that makes them combustible is a hypothesis. Suppose you notice that all the items on 
the "will burn" list are cylindrical in shape and therefore, you hypothesize that "cylindrical objects burn". 
The fourth step in the scientific method is to test your hypothesis. To test this hypothesis, you go out 
and collect a group of objects that are cylindrical including iron pipes, soda bottles, broom handles, and 
tin cans. When these cylindrical objects are placed in the fire and most of them don't burn, you realize 
your hypothesis is not supported by these new observations. The new observations are the test, and your 
hypothesis has failed the test. When the new observations fail to support your hypothesis, you reject your 
original hypothesis, add your new data to (Table 1.3), and make a new hypothesis based on the updated 
observations list. In the schematic diagram of the scientific method, a failed test returns the scientist to 
step 3, make a new hypothesis. 



11 www.ckl2.org 



Table 1.3: 



Will Burn Won't Burn 



tree limbs rocks 

chair legs bricks 

pencils marbles 

baseball bat hubcaps 

broom handle iron pipes 

soda bottles 
tin cans 



Suppose your new hypothesis is "wooden objects burn." You will find this hypothesis more satisfactory 
since all the wooden object you try will burn. Your confidence will grow that you have discovered a "law 
of nature." Even with your somewhat successful theory, you might be ignoring a large stack of old car tires, 
objects made of fabric or paper, or perhaps containers of petroleum. You can see that even though you 
are quite satisfied with your theory because it does the job you want it to do, you actually do not have a 
complete statement on the property of objects that make them burn. So it is with science. 

You can see from this example that the "solution" does not become what we think of as a "fact," but rather 
becomes a tentatively accepted theory which must undergo continuous testing and perhaps 
adjustment. No matter how long a tentative explanation has been accepted, it can be discarded at any 
time if contradictory observations are found. As long as the theory is consistent with all observations, 
scientists will continue to use it. When a theory is contradicted by observations, it is discarded and 
replaced. Even though the terms hypothesis, theory, and fact are used somewhat carelessly at times, a 
theory will continue to be used while it is useful and will be called into question when contradictory 
evidence is found. Theories never become facts. 

There is a common generalization about theories, which says that "theories are much easier to disprove 
than to prove." The common example given is a hypothesis that "all swans are white." You may observe a 
thousand white swans and every observation of a white swan supports your hypothesis, but it only takes 
a single observation of a black swan to disprove the hypothesis. To be an acceptable scientific hypothesis, 
observations that disprove the hypothesis must be possible. That is, if every conceivable observation 
supports the hypothesis, then it is not an acceptable scientific hypothesis. To be a scientific hypothesis, it 
must be possible to refute the concept. 



Some Basic Terminology 

• A hypothesis is a guess that is made early in the process of trying to explain some set of observations. 
There are scientists who object to calling a hypothesis a "guess". The primary basis for the objection 
is that someone who has studied the subject under consideration would make a much better guess 
than someone who was completely ignorant of the field. Perhaps we should say that a hypothesis is 
an "educated guess." 

• A theory is an explanation that stands up to everyday use in explaining a set of observations. A 
theory is not proven and is not a "fact." A scientific theory must be falsifiable in order to be accepted 
as a theory. 

• A law describes an observable relationship, that is, observations that occur with a predictable rela- 
tionship to each other. It is only after experience shows the law to be valid that it is incorporated 
into the field of knowledge. 

www.ckl2.org 12 



The Scientific Revolution 

The explosion of achievement in the last 160 years was produced by using a new method for learning 
about nature. This sudden and massive achievement in understanding nature is called the Scientific 
Revolution and was produced using the scientific method. 

The British historian, Herbert Butterfield, wrote a book called The Origins of Modern Science. In the 
preface to the book, Butterfield wrote: 

The Revolution in science overturned the authority of not only the Middle Ages but of the ancient world . 
. . it ended not only in the eclipse of scholastic philosophy but in the destruction of Aristotelian physics. 
The Scientific Revolution outshines everything since the rise of Christianity and reduces the Renaissance 
and Reformation to the rank of mere episodes . . . 

The beginning of the 17 th century is known for the drastic changes that occurred in the European approach 
to science during that period and is known as the Scientific Revolution. This term refers to a completely 
new era of academic thought in which medieval philosophy was abandoned in favor of innovative methods 
offered by Galileo and Newton. 

Science is best defined as a careful, disciplined, logical search for knowledge about any and all aspects of 
the universe, obtained by examination of the best available evidence and always subject to correction and 

improvement upon discovery of better evidence. What's left is magic. And it doesn't work. James 

Randi. 

What is an Experiment 

The scientific method requires than observations be made. Sometimes, the phenomenon we wish to observe 
does not occur in nature or if it does, it is inconvenient for us to observe. Therefore, it is more successful 
for us to cause the phenomenon to occur at a time and place of our choosing. When we arrange for the 
phenomenon to occur at our convenience, we can have all our measuring instruments present and handy to 
help us make observations, and we can control other variables. Causing a phenomenon to occur when and 
where we want it and under the conditions we want is called an experiment. When scientists conduct 
experiments, they are usually seeking new information or trying to verify someone else's data. Classroom 
experiments often demonstrate and verify information that is already known but new to the student. When 
doing an experiment, it is important to set up the experiment so that relationships can be seen clearly. 
This requires what are called experimental controls. 

Experimental Controls 

Suppose a scientist, while walking along the beach on a very cold day following a rainstorm, observed two 
pools of water in bowl shaped rocks near each other. One of the pools was partially covered with ice while 
the other pool had no ice on it. The unfrozen pool seemed to be formed from seawater splashing up on 
the rock from the surf, but the other pool was too high for sea water to splash in, so it was more likely to 
have been formed from rainwater. 

The scientist wondered why one pool was partially frozen and not the other since both pools were at the 
same temperature. By tasting the water (not a good idea), the scientist determined that the unfrozen pool 
tasted saltier than the partially frozen one. The scientist thought perhaps salt water had a lower freezing 
point that fresh water and she decided to go home and try an experiment to see if this were true. So far, 
the scientist has identified a question, gathered a small amount of data, and suggested a hypothesis. In 
order to test this hypothesis, the scientist will conduct an experiment during which she can make accurate 
observations. 

13 www.ckl2.org 





For the experiment, the scientist prepared two identical containers of fresh water and added some salt to 
one of them (Figure ??). A thermometer was placed in each liquid and these were put in a freezer. The 
scientist then observed the conditions and temperatures of the two liquids at regular intervals. 

The Temperature and Condition of Fresh Water in a Freezer 



Time (minu 


tes) 5 


10 15 


20 


25 


30 


Temp., C 


25 20 


15 10 


5 





-5 


Condition 
Temperature 


Liquid Liquid 
e and Condition of Salt Water 


Liquid Liquid 
in a Freezer 


Liquid 


Frozen 


Frozen 


Time (minutes) 5 


10 15 


20 


25 


30 


Temp., C 


25 20 


15 10 


5 





-5 


Condition 


Liquid Liquid 


Liquid Liquid 


Liquid 


Liquid 


Frozen 



The scientist found support for the hypothesis from this experiment; fresh water freezes at a higher tem- 
perature than salt water. Much more support would be needed before the scientist would be confident of 
this hypothesis. Perhaps she would ask other scientists to verify the work. 

In the scientist's experiment, it was necessary that she freeze the salt water and fresh water under exactly 
the same conditions. Why? The scientist was testing whether or not the presence of salt in water would 
alter its freezing point. It is known that changing air pressure will alter the freezing point of water. In 
order to conclude that the presence of the salt was what caused the change in freezing point, all other 
conditions had to be identical. The presence of the salt is called the experimental variable because 
it is the only thing allowed to change in the two trials. The fresh water part of the experiment is called 
the experimental control. In an experiment, there may be only one variable and the purpose of the 
control is to guarantee that there is only one variable. The "control" is identical to the "test" except for 
the experimental variable. Unless experiments are controlled, the results are not valid. 

Suppose you wish to determine which brand of microwave popcorn leaves the fewest unpopped kernels. 
You will need a supply of various brands of microwave popcorn to test and you will need a microwave 
oven. If you used different brands of microwave ovens with different brands of popcorn, the percentage 
of unpopped kernels could be caused by the different brands of popcorn, but it could also be caused by 
the different brands of ovens. Under such circumstances, the experimenter would not be able to conclude 
confidently whether the popcorn or the oven caused the difference. To eliminate this problem, you must 
use the same microwave oven for every test. By using the same microwave oven, you control the number 
of variables in the experiment. 

What if you allowed the different samples of popcorn to be cooked at different temperatures? What if 
you allowed longer heating periods? In order to reasonably conclude that the change in one variable was 



www.ckl2.org 



14 



caused by the change in another specific variable, there must be no other variables in the experiment. All 
other variables must be kept constant. 

Errors in the Use of the Scientific Method 

The scientific method requires the observation of nature and correspondence between the suggested ex- 
planation and the observations. That is, the hypothesis must explain all the observations. Therefore, 
the scientific method can only work properly when the data (observation list) is not biased. There are 
several ways in which a biased set of data can be produced. It is always possible for anyone to make 
an error in observation. A balance can be misread or numbers can be transposed when written down. 
That is one of the reasons that experiments are run several times and the observations made over and 
over again. It is also possible that an unrecognized error is present and produces the same error in every 
experiment. For example, a scientist may be attempting to test normal rainwater, but unknown to him 
a nearby factory is sending soluble substances out of their smoke stack and the material is contaminating 
the scientist's samples. In such a case, the scientist's samples would yield false observations for normal 
rainwater. Other scientists reproducing the experiment would collect uncontaminated samples and find 
different results. Multiple testing of the experiment would determine which set of data was flawed. Failing 
to apply appropriate experimental controls would certainly bias data. 

There are also dishonest mistakes that occur when the experimenter collects only supporting data and 
excludes contradictory observations. There have even been scientists who faked observations to provide 
support for his/her hypothesis. The attractions of fame and fortune can be hard to resist. For all these 
reasons, the scientific method requires that experimental results be published and the experiment be 
repeated by other scientists. 

Lesson Summary 

• Before the development of the scientific method, mankind made only slight achievements in the areas 
of transportation, communication and medicine. 

• Use of the scientific method allowed mankind to make significant achievements in transportation, 
communication, and medicine. 

• The scientific method has been much more successful than the methods of superstition, opinion, and 
authority. 

• The steps in the scientific method are: 

1. Identify the problem. 

2. Gather data (make observations). 

3. Suggest a hypothesis. 

4. Test the hypothesis (experiment). 

5. Continue testing or reject the hypothesis and make a new one. 

• Experimental controls are used to make sure that the only variables in an experiment are the ones 
being tested. 

Review Questions 

Use the following paragraph to answer questions 1 and 2. In 1928, Sir Alexander Fleming was studying 
Staphylococcus bacteria growing in culture dishes. He noticed that a mold called Penicillium was also 
growing in some of the dishes. In Figure ??, Petri dish A represents a dish containing only Staphylococcus 

15 www.ckl2.org 



bacteria. The red dots in dish B represent Penicillium colonies. Fleming noticed that a clear area existed 
around the mold because all the bacteria grown in this area had died. In the culture dishes without the 
mold, no clear areas were present. Fleming suggested that the mold was producing a chemical that killed 
the bacteria. He decided to isolate this substance and test it to see if it would kill bacteria. Fleming grew 
some Penicillium mold in a nutrient broth. After the mold grew in the broth, he removed all the mold 
from the broth and added the broth to a culture of bacteria. All the bacteria died. 




B 



1. Which of the following statements is a reasonable expression of Fleming's hypothesis? 

(a) Nutrient broth kills bacteria. 

(b) There are clear areas around the Penicillium mold where Staphylococcus doesn't grow. 

(c) Mold kills bacteria. 

(d) Penicillium mold produces a substance that kills Staphylococcus. 

(e) Without mold in the culture dish, there were no clear areas in the bacteria. 

2. Fleming grew Penicillium in broth, then removed the Penicillium and poured the broth into culture 
dishes containing bacteria to see if the broth would kill the bacteria. What step in the scientific 
method does this represent? 

(a) Collecting and organizing data 

(b) Making a hypothesis 

(c) Testing a hypothesis by experiment 

(d) Rejecting the old hypothesis and making a new one 

(e) None of these 

3. A scientific investigation is NOT valid unless every step in the scientific method is present and carried 
out in the exact order listed in this chapter. 

(a) True 

(b) False 

4. Which of the following words is closest to the same meaning as hypothesis? 

(a) fact 

(b) law 

(c) formula 

(d) suggestion 

(e) conclusion 

5. Why do scientists sometimes discard theories? 

(a) the steps in the scientific method were not followed in order 

(b) public opinion disagrees with the theory 

(c) the theory is opposed by the church 

(d) contradictory observations are found 

(e) congress voted against it 



www.ckl2.org 16 



Gary noticed that two plants which his mother planted on the same day that were the same size when 
planted were different in size after three weeks. Since the larger plant was in the full sun all day and the 
smaller plant was in the shade of a tree most of the day, Gary believed the sunshine was responsible for 
the difference in the plant sizes. In order to test this, Gary bought ten small plants of the size and type. 
He made sure they had the same size and type of pot. He also made sure they have the same amount and 
type of soil. Then Gary built a frame to hold a canvas roof over five of the plants while the other five 
were nearby but out in the sun. Gary was careful to make sure that each plant received exactly the same 
amount of water and plant food every day. 

6. Which of the following is a reasonable statement of Gary's hypothesis? 

(a) Different plants have different characteristics. 

(b) Plants that get more sunshine grow larger than plants that get less sunshine. 

(c) Plants that grow in the shade grow larger. 

(d) Plants that don't receive water will die. 

(e) Plants that receive the same amount of water and plant food will grow the same amount. 

7. What scientific reason might Gary have for insisting that the container size for the all plants be the 
same? 

(a) Gary wanted to determine if the size of the container would affect the plant growth. 

(b) Gary wanted to make sure the size of the container did not affect differential plant growth in 
his experiment. 

(c) Gary want to control how much plant food his plants received. 

(d) Gary wanted his garden to look organized. 

(e) There is no possible scientific reason for having the same size containers. 

8. What scientific reason might Gary have for insisting that all plants receive the same amount of water 
everyday? 

(a) Gary wanted to test the effect of shade on plant growth and therefore, he wanted to have no 
variables other than the amount of sunshine on the plants. 

(b) Gary wanted to test the effect of the amount of water on plant growth. 

(c) Gary's hypothesis was that water quality was affecting plant growth. 

(d) Gary was conserving water. 

(e) There is no possible scientific reason for having the same amount of water for each plant every 
day. 

9. What was the variable being tested in Gary's experiment? 

(a) the amount of water 

(b) the amount of plant food 

(c) the amount of soil 

(d) the amount of sunshine 

(e) the type of soil 

10. Which of the following factors may be varying in Gary's experimental setup that he did not control? 

(a) individual plant variation 

(b) soil temperature due to different colors of containers 

(c) water loss due to evaporation from the soil 

(d) the effect of insects which may attack one set of plants but not the other 

(e) All of the above are possible factors that Gary did not control 

11. When a mosquito sucks blood from its host, it penetrates the skin with its sharp beak and injects 
an anti-coagulant so the blood will not clot. It then sucks some blood and removes its beak. If the 

17 www.ckl2.org 



mosquito carries disease-causing microorganisms, it injects these into its host along with the anti- 
coagulant. It was assumed for a long time that the virus of typhus was injected by the louse when 
sucking blood in a manner similar to the mosquito. But apparently this is not so. The infection is 
not in the saliva of the louse, but in the feces. The disease is thought to be spread when the louse 
feces come in contact with scratches or bite wounds in the host's skin. A test of this was carried out 
in 1922 when two workers fed infected lice on a monkey taking great care that no louse feces came 
into contact with the monkey. After two weeks, the monkey had NOT become ill with typhus. The 
workers then injected the monkey with typhus and it became ill within a few days. Why did the 
workers inject the monkey with typhus near the end of the experiment? 

(a) to prove that the lice carried the typhus virus 

(b) to prove the monkey was similar to man 

(c) to prove that the monkey was not immune to typhus 

(d) to prove that mosquitoes were not carriers of typhus 

(e) the workers were mean 

12. Eijkman fed a group of chickens exclusively on rice whose seed coat had been removed (polished 
rice or white rice). The chickens all developed polyneuritis (a disease of chickens) and died. He fed 
another group of chickens unpolished rice (rice that still had its seed coat). Not a single one of them 
contracted polyneuritis. He then gathered the polishings from rice (the seed coats that had been 
removed) and fed the polishings to other chickens that were sick with polyneuritis. In a short time, 
the birds all recovered. Eijkman had accurately traced the cause of polyneuritis to a faulty diet. 
For the first time in history, a food deficiency disease had been produced and cured experimentally. 
Which of the following is a reasonable statement of Eijkman's hypothesis? 

(a) Polyneuritis is a fatal disease for chickens. 

(b) White rice carries a virus for the disease polyneuritis. 

(c) Unpolished rice does not carry the polyneuritis virus. 

(d) The rice seed coat contains a nutrient that provides protection for chickens against polyneuritis. 

(e) None of these is a reasonable statement of Eijkman's hypothesis. 

Questions 12, 13, and 14 relate to the following paragraphs. 

Scientist A noticed that in a certain forest area, the only animals inhabiting the region were giraffes. 
He also noticed that the only food available for the animals was on fairly tall trees and as the summer 
progressed, the animals ate the leaves high and higher on the trees. The scientist suggested that these 
animals were originally like all other animals but generations of animals stretching their necks to reach 
higher up the trees for food, caused the species to grow very long necks. 

Scientist B conducted experiments and observed that stretching muscles does NOT cause bones to grow 
longer nor change the DNA of animals so that longer muscles would be passed on to the next generation. 
Scientist B, therefore, discarded Scientist A's suggested answer as to why all the animals living in the area 
had long necks. Scientist B suggested instead that originally many different types of animals including 
giraffes had lived in the region but only the giraffes could survive when the only food was high in the trees, 
and so all the other species had left the area. 

12. Which of the following statements is an interpretation, rather than an observation? 

A. The only animals living in the area were giraffes. 

B. The only available food was on tall trees. 

C. Animals which constantly stretch their necks will grow longer necks. 

D. A, B, and C are all interpretations. 

E. A, B, and C are all observations. 

www.ckl2.org 18 



13. Scientist A's hypothesis was that 

A. the only animals living in the area were giraffes. 

B. the only available food was on tall trees. 

C. animals which constantly stretch their necks will grow longer necks. 

D. the animals which possess the best characteristics for living in an area, will be the predominant species. 

E. None of the above are reasonable statements of Scientist A's hypothesis. 

14. Scientist A's hypothesis being discarded is 

A. evidence that the scientific method doesn't always work. 

B. a result achieved without use of the scientific method. 

C: an example of what happened before the scientific method was invented. 

D. an example of the normal functioning of the scientific method. 

E. an unusual case. 

15. When a theory has been known for a long time, it becomes a law. 

A. True 

B. False 

16. During Pasteur's time, anthrax was a widespread and disastrous disease for livestock. Many people 
whose livlihood was raising livestock lost large portions of their herds to this disease. Around 1876, a horse 
doctor in eastern France named Louvrier, claimed to have intvented a cure for anthrax. The influential 
men of the community supported Louvrier's claim to have cured hundreds of cows of anthrax. Pasteur 
went to Louvrier's hometown to evaluate the cure. The cure was explained to Pasteur as a multi-step 
process during which: 1) the cow was rubbed vigorously to make her as hot as possible; 2) long gashes 
were cut into the cows skin and turpentine was poured into the cuts; 3) an inch-thick coating of cow 
manure mixed with hot vinegar was plastered onto the cow and the cow was completely wrapped in a 
cloth. Since some cows recover from anthrax with no treatment, performing the cure on a single cow would 
not be conclusive, so Pasteur proposed an experiment to test Louvrier's cure. Four healthy cows were to 
be injected with anthrax microbes, and after the cows became ill, Louvrier would pick two of the cows 
(A and B) and perform his cure on them while the other two cows (C and D) would be left untreated. 
The experiment was performed and after a few days, one of the untreated cows died and one of them got 
better. Of the cows treated by Louvrier's cure, one cow died and one got better. In this experiment, what 
was the purpose of infecting cows C and D? 

A. So that Louvrier would have more than two cows to choose from. 

B. To make sure the injection actually contained anthrax. 

C. To serve as experimental controls (a comparison of treated to untreated cows). 

D. To kill as many cows as possible. 

17. A hypothesis is 

A. a description of a consistent pattern in obervations. 

B. an observation that remains constant. 

C. a theory that has been proven. 

D. a tentative explanation for a phenomenon. 

19 www.ckl2.org 



18. A scientific law is 

A. a description of a consistent pattern in obervations. 

B. an observation that remains constant. 

C. a theory that has been proven. 

D. a tentative explanation for a phenomenon. 

19. A number of people became ill after eating oysters in a restaurant. Which of the following statements 
is a hypothesis about this occcurence? 

A. Everyone who ate oysters got sick. 

B. People got sick whether the oysters they ate were raw or cooked. 

C. Symptoms included nausea and dizziness. 

D. The cook felt really bad about it. 

E. Bacteria in the oysters may have caused the illness. 

20. Which statement best describes the reason for using experimental controls? 

A. Experimental controls eliminate the need for large sample sizes. 

B. Experimental controls eliminate the need for statistical tests. 

C. Experimental controls reduce the number of measurements needed. 

D. Experimental controls allow comparison between groups that are different in only one independent 

variable. 

21. A student decides to set up an experiment to determine the relationship between the growth rate of 
plants and the presence of detergent in the soil. He sets up 10 seed pots. In five of the seed pots, he 
mixes a precise amount of detergent with the soil and the other five seed pots have no detergent in the 
soil. The 5 seed pots with detergent are placed in the sun and the five seed pots with no detergent are 
placed in the shade. All 10 seed pots receive the same amount of water and the same number and type of 
seeds. He grows the plants for two months and charts the growth every two days. What is wrong with his 
experiment? 

A. The student has too few pots. 

B. The student has two independent variables. 

C. The student has two dependent variables. 

D. The student has no experimental control on the soil. 

22. A scientist plants two rows of corn for experimentation. She puts fertilizer on row 1 but does not put 
fertilizer on row 2. Both rows receive the same amount of sun and water. She checks the growth of the 
corn over the course of five months. What is acting as the control in this experiment? 

A. Corn without fertilizer. 

B. Corn with fertilizer. 

C. Amount of water. 

D. Height of corn plants. 

23. If you have a control group for your experiment, which of the following is true? 
www.ckl2.org 20 



A. There can be more than one difference between the control group and the test group, but not more 

three differences or else the experiment is invalid. 

B. The control group and the test group may have many differences between them. 

C. The control group must be identical to the test group except for one variable. 

D. None of these are true. 

24. If the hypothesis is rejected by the experiment, then: 

A. the experiment may have been a success. 

B. the experiment was a failure. 

C. the experiment was poorly designed. 

D. the experiment didn't follow the scientific method. 

25. A well-substantiated explanation of an aspect of the natural world is a: 

A. theory. 

B. law. 

C. hypothesis. 

D. None of these. 

Further Reading / Supplemental Links 

• http : //learner . org/resources/series61 . html 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. The website has two 
videos that apply to this lesson. One is a video called The World of Chemistry that relates chemistry 
to other sciences and daily life. Another video called Thinking Like Scientists relates to the scientific 
method. The audience on the video is young children but the ideas are full grown. 

• Website with lessons, worksheets, and quizzes on various high school chemistry topics. Lesson 1-2 is on 
the scientific method, http://www.fordhamprep.org/gcurran/sho/sho/lessons/lessonl2.htm 

• Website of the James Randi Foundation. James Randi is a staunch opponent of fake science, http : 
//www . randi . org/site/ 

• Websites dealing with the history of the scientific method, http : //www. historyguide . org/earlymod/ 
lecturelOc . html http : //www . history . boisestate . edu/WESTCIV/science/ 

Vocabulary 

hypothesis A proposal intended to explain a set of observations. 

theory A hypothesis that has been supported with repeated testing. 

law A relationship that exists between specific observations. 

experiment The act of conducting a controlled test or observations. 

21 www.ckl2.org 



scientific method A method of investigation involving observation to generate and test hypotheses and 
theories. 

superstition An irrational belief that an object, action, or circumstance not logically related to an event 
influences its outcome. 

Labs and Demonstrations for The Scientific Method 
Teacher's Resource Page for Candle Observation 

Investigation and Experimentation Objectives 




In this activity, the student will be making and recording observations. 

Safety Issues 

If students are allowed to light their own candles, they should be instructed to strike matches on the 
striker pad in a direction away from the body such that potential flying pieces of the burning match head 
move away from the body. Extinguished matches should be held until cool, and then placed in solid 
waste containers (wastebasket). Students should be reminded that during any lab involving an open flame 
(candles, Bunsen burners, etc.) long hair must be restrained behind the head so that it does not fall 
past the face when looking down. Students should be instructed not to handle candles once they are lit. 
Dripping hot wax can be painful. 

Observation List 

The candle is cylinder in shape 1 and has a diameter 2 of about 2 cm. The length 3 of the candle was initially 
about 18 cm and changed slowly during observation 4 , decreasing about 4 mm in 20 minutes. 

The candle is made of a translucent, 5 white 6 solid 7 which has a slight odor 8 and no taste. 9 It is soft enough 
to be scratched with a fingernail. 10 There is a wick which extends from the top to bottom of the candle 
along its central axis 11 and protrudes 12 about 10 mm above the top of the candle. The wick is made of 
three strands 13 of string braided 14 together. 

The candle is lit by holding a source of flame close to the wick for a few seconds. 15 Thereafter, the source 
of the flame can be removed and the flame sustains itself 16 at the wick. 

The burning candle makes no sound. 17 While burning, the body of the candle remains cool to the touch 18 
except near the top. 19 Within about 5 mm from the top of the candle, it is warm 20 but not hot, and 
sufficiently soft to mold 21 easily. 

The flame flickers 22 in response to air currents and tends to become smoky 23 while flickering. In the absence 
of air currents, the flame is in the form shown in the picture 24 although it retains some movement 25 at all 
times. 

The flame begins 26 about 4 mm above the top of the candle, and at its base, the flame has a blue tint. 27 
Immediately around the wick in a region about 5 mm wide 28 and extending about 8 mm above 29 the top 
of the wick, the flame is dark. 30 This dark region is roughly conical 31 in shape. 

Around this dark zone and extending about 5 mm above the dark zone is a region which emits yellow 
light, 32 bright 33 but not blinding. The flame has rather sharply defined sides 34 but a ragged 35 top. 

The wick is white 36 where it emerges from the candle, but from the base of the flame to the end of wick, it 
is black, 37 appearing burnt except for the last 2 mm where it glows red. 38 The wick curls 39 over about 4 mm 

www.ckl2.org 22 



from its end. As the candle becomes shorter, the wick shortens 40 too, so as to extend roughly a constant 
distance above the top of the candle. 

Heat is emitted 41 by the flame, enough so that it becomes uncomfortable in a few seconds to hold ones' 
fingers near the flame. 

The top of a quietly burning candle becomes wet 42 with a colorless 43 liquid and becomes bowl-shaped. 44 
Sometimes, the liquid in the bowl drains 45 down the side of the candle, cools, gradually solidifying 46 and 
attaching itself to the candle. 47 

Under quiet conditions, a stable 48 pool of clear liquid remains in the bowl-shaped top of the candle. The 
liquid rises slightly around the wick, wetting 50 the base of the wick as high as the base 51 of the flame. 

Candle Observation 

Materials 




Each student or pair of students is given a candle to observe. The candles should be between ^ an d 3 inch 
in diameter (so their length will significantly shorten during the period) and mounted securely in a candle 
holder (a jar lid will work fine). The teacher should light the candles, give appropriate instructions about 
hair (it burns), and other safety issues. The student should be instructed to make as many observations 
about the burning candle as they can in the allotted time. The teacher should encourage estimated 
quantitative observations. 

After the observation period, the teacher can ask for observations from the class and get as many as possible 
on the board. Further observations can be added from the list above. It is useful for students to recognize 
that there are many more observations about a simple system than they may have imagined. 

Teacher's Resource Page for DAZOO 

Investigation and Experimentation Objectives 




In this activity, the student will use critical thinking to note evidence and logically draw a conclusion. 



23 www.ckl2.org 



Table 1.4: Answers for DAZOO 



Question Number 



Answer 



Reason 



4 
5 

6 

7. 
8 
9 
10 



11. 

12. 

13. 
14. 
15. 
16. 



ZAM 

NOOT 

NIX 

GOBBIE 
BOBO 

YATZ 

BOBO 

YATZ 

CLINT 

ARDZU 



Girl 

SLIP 

BOBO 
YATZ 
WHEE 
NIX 



It is a family and they are outside 
the cages. 

It is a single male outside the 
cages. 

It is the largest number of indi- 
viduals in a cage. 
They have eight legs. 
There is no large circle in the 
cage. 

There is no large rectangle in the 
cage. 

There are two small circles in the 
cage. 

There are three small rectangles 
in the cage. 

There are no small circles or rect- 
angles in the cage. 
Circle in circle - mother is preg- 
nant - can't be SLIP because the 
child is inside of a male. 
The baby is a circle - which rep- 
resents female. 

There is a small BOBO rectangle 
inside the large SLIP rectangle. 
The crocodile ate a small giraffe. 
They have wings. 
They have no legs. 
They live between YATZ and 
CLINT. 



Observation Game: DAZOO 



SET-UP: Teacher prints out a copy of the game image and a set of questions for each student. With a 
little thought, most students can answer the questions without teacher input. 



www.ckl2.org 



24 



GOBBIE 




BOBO 



O 



WHEE 




ARDZU 




DAZOO 



NOOT 



D 



ZAM 



D 








YATZ 




NIX 



:: 



u 




CLINT 




SLIP 




x n x 

xVx 




Only one family and the zookeeper are at the zoo today. The zookeeper is a single male, but the family 
visiting the zoo has both males and females in the family. The family groupings of the zoo families have 
been given surnames and are identified by the surnames in the diagram. Try to answer the questions below. 

1 . What is the name of the family visiting the zoo? 

2. What is the name of the zookeeper? 

3. Which family in the zoo has the most members? 

4. What is the name of the family of spiders? 

5. Which mother is away at the hospital? 

6. Which family has no father? 

7. Which family has two daughters? 

8. Which family has three sons? 

9. Mr. and Mrs. Elephant have no children. Which family are they? 
10. What will be the last name of the baby tiger when it is born? 



25 



www.ckl2.org 



11. Will the baby tiger be a boy or a girl? 

12. Mr. Crocodile has swallowed the giraffes son. Which is the family of crocodiles? 

13. Which is the family of giraffes? 

14. Which is the family of pelicans? 

15. Which is the family of snakes? 

16. The aardvarks live between the pelicans and elephants. Which is the family of aardvarks? 



The Seven-of-Diamonds Game 




Investigation and Experimentation Objectives 



In this activity, the student will make and record observations, generate a hypothesis, and test the hy- 
pothesis against further experimental observations. If the hypothesis fails, the student will add the new 
observations to his list and create another hypothesis. If the hypothesis succeeds, the student "wins" the 
game. 

How to Play 

This is the easiest of the observation and hypothesis games. To play the game, the teacher must select and 
instruct an assistant to play the role of "psychic". The teacher draws the seven-of-diamonds set up on the 
board as shown below. 



!♦♦ 



♦ ♦: 



\ ♦ 


u ♦ 


!♦ ♦ 

♦ 


♦ \ 


♦ ♦< 


♦ ♦: 









♦ ♦ 



To begin the game, the "psychic" is sent out of the room. While the psychic is out of the room, the 
students select one of the cards and inform the teacher which card. Then the psychic is called back into 
the room and the teacher points to one of the cards and asks the psychic, "Is this the card?" The psychic 
responds either "yes" or "no" and the process continues until the teacher points to the correct card and the 
psychic correctly identifies the card as the one the students had selected. The students are to observe the 
game and after each trial run, make hypotheesis about how the trick is being done. They can do this as 
individuals or in groups, the game can be played over and over until at least one student or group figures 
out how the trick is being done. When a student or group thiks they know the trick, they can go out of 
the room with the psychic and then play the role of the psychic when they return. If they can correctly 



www.ckl2.org 



26 



identify the selected card, they win the game. The game can continue until more students figure it out or 
the winners can explain the trick to those who didn't figure it out. 



!♦♦ 



♦ ♦; 



I ♦ 


!♦ ♦ 


!♦ ♦ 




♦♦♦ 


♦ 


♦ i 


z t\ 


♦ ♦; 



♦ 




♦ ♦ 



THE TRICK 

The layout of the cards on the board and the spots on the seven-of-diamonds exactly correspond. The 
psychic will not know the correct card until the teacher points to the seven-of-diamonds and asks, "Is this 
the card?" When the teacher points to the seven-of-diamonds, he/she points to the spot on the card that 
corresponds to the card selected. In the picture above, the teacher is pointing to the spot on the seven- 
of-diamonds that corresponds to the position of the "6" in the layout. Therefore, the six-of-diamonds is 
the selected card for this trial. The psychic continues to say "no" until the teacher points to the six-of- 
diamonds and then says "yes." You should note, it is not possible to have the psychic identify the correct 
card on the first try unless the selected card is the seven-of-diamonds. If the seven is the selected card, 
the teacher can point to the seven first and point to the center position. The psychic must be alert to get 
this one. The teacher can vary the sequence of asking so that sometimes, the selected card is pointed to 
on the second try or the fourth try, and so forth. 



"This" or "That" Psychic Game 



I i ) 



Investigation and Experimentation Objectives 



In this activity, the student will make and record observations, generate a hypothesis, and test the hy- 
pothesis against further experimental observations. If the hypothesis fails, the student will add the new 
observations to his list and create another hypothesis. If the hypothesis succeeds, the student "wins" the 
game. 

How to Play 

This is the most difficult of these observation/hypothesis games for the students to figure out. The teacher 
draws 3 columns of 3 squares each on the board as shown at right. Once again, the teacher needs an 
assistant to act as psychic. Secretly, the teacher and the "psychic" conspire and assign the two outside 
columns to be called "this" columns and the middle column to be called a "that" column. 



27 



www.ckl2.org 



As usual, the psychic leaves the room and the students select one of the squares to be "psychically" 
identified. The psychic is called back into the room and the teacher proceeds to point at various squares 
and ask the psychic, "Is it this one?, or "Is it that one?" The code known only to the teacher and the 
psychic is that if the teacher uses the correct name of the column when inquiring about a square, the 
psychic answers "no." If the teacher uses the incorrect name of the column when inquiring, the psychic 
replies "yes." 




□ □□ 



This Column That Column This Column 





Is it this one? (Correct title) NO 




Ts it this one? (Correct title) NO 




Is it this one? (Incorrect title) YES 



One of the things that make this game so difficult is that the teacher can ask about the correct square on 
the first try. On the very first trial, the teacher can point to a square in the middle column (the "that" 



www.ckl2.org 



28 



column) and ask, "Is it this one?" and the psychic replies "yes." 

1.2 Chemistry in History 

Lesson Objectives 



Give a brief history of how chemistry began. 

State the Law of Conservation of Mass. 

Explain the concept of a model, and create simple models from observations. 



Introduction 



During medieval times, a group of people known as alchemists began looking for ways to transform common 
metals, such as lead, copper and iron, into gold (Figure ??). Can you imagine how much money you would 
make if you could go to the store, buy some iron nails, and turn them into gold? You'd be rich in no time! 

Alchemists experimented with many different kinds of chemicals, searching for what they termed the 
"philosopher's stone" - a legendary substance that was necessary for the transformation of common metals 
into gold. We now know that there is no such thing as a "philosopher's stone," nor is there any chemical 
reaction that creates gold from another metal. We know this because we now have a much better under- 
standing of the matter in our universe. Nevertheless, it was thanks to those early alchemists that people 
became interested in chemistry in the first place. 

29 www.ckl2.org 




"Chemistry" was Derived from an Arabic Word 

When we speak of "chemistry," we refer to the modern, scientific study of matter and the changes that it 
undergoes. Still, it's no coincidence that the word "chemistry," looks a lot like the word "alchemy." Early 
alchemists were commonly known as 'chemists,' and over time, people started referring to their work, 
particularly the more legitimate forms of it, as chemistry. In many ways, it's appropriate that our word for 
the present-day study of matter comes from the early practice of alchemy because a lot of the techniques 
and equipment fundamental to modern chemistry were actually developed by early alchemists. 

The origin of the word "alchemy" is something of a mystery. Certainly, early Europeans borrowed 
"alchemy" from the Arabic word "al-kimia, " meaning "the art of transformation" (of course, the trans- 
formation that alchemists were primarily concerned with involved the creation of gold). Most of what we 
know today about early alchemy is based on translations of Arabic documents. That's because Muslim 
alchemists were some of the first to keep careful notes about their experiments. 

Even though our earliest records of alchemy come from the Arab Empire, some scholars believe that Arabs 
adopted alchemy and the word "al-kimia" from the Greeks around 650 AD. The Greeks, in turn, may have 
learned of alchemy from the Egyptians. Khem was an ancient name for Egypt, and Egyptians were known, 
in early history, as masters of the art of working with gold. It's very likely that "al-kimia" is actually a 
distorted version of the word "al-kimiya," meaning "the art of the land of Khem," or the art of Egypt. 



www.ckl2.org 



30 



The Origins of Chemistry was Multi- Cultural 

While the word "chemistry" may have its roots in Egypt, chemical experimentation seems to have been 
prevalent all over the world, even as early as the 5 century BC. In China, the goal of the early alchemists 
was largely to find "the elixir of life" - a potion that could cure all diseases and prevent death. Ironically, 
many of these early elixirs involved mixtures of mercury and arsenic salts, both of which are extremely 
poisonous. In fact, it's rumored that several Chinese emperors actually died after drinking "elixirs of life" 
(Figure 1.10). Despite never finding a magical potion that could cure all diseases, early Chinese chemists 
did discover many new chemicals and chemical reactions, including those used in fireworks and gunpowder. 




Figure 1.10: Jiajing Emperor, rumored to have died after drinking poison he believed to be "the elixir of 
life." 

Like the Chinese, alchemists from India were interested in some of the medical benefits of different chem- 
icals. In addition to medicine, Indian alchemists were fascinated by metals and metallurgy. Early Indian 
writings contain methods for extracting and purifying metals like silver, gold and tin from ores that were 
mined out of the ground. Moreover, it was alchemists from the Indian subcontinent who first realized that 
by mixing molten metals with other chemicals they could produce materials that had new and beneficial 
properties. For example, Wootz steel (also known as Damascus steel) was a substance discovered in Sri 
Lanka around 300 AD. It was made by mixing just the right amounts of molten iron, glass and charcoal, but 
it became famous because it could be used to produce swords, legendary for their sharpness and strength 
in battle. 

Around the same time that Wootz steel was being developed in Sri Lanka, Egyptian and Greek alchemists 
were beginning to experiment as well. Much of the alchemy in this part of the world involved work 
with colors and dyes and, of course, the transformation of common metals into gold and silver. Greek 
philosophers like Plato and Aristotle, however, were also responsible for the important suggestion that 



31 



www.ckl2.org 



the universe could be explained by unified natural laws. As you will discover in the next section, modern 
chemistry relies on the study of these universal laws. Of course, modern scientific laws are slightly different 
than the laws that either Plato or Aristotle had in mind. In general, scientific laws are determined by 
careful experimentation and observation, whereas the early Greeks believed that their "natural laws" 
could be deduced through philosophy. 

Medieval Europeans were similarly fascinated by alchemy. Unfortunately, many alchemists in Europe 
borrowed ideas from the more mystical of the Arabian alchemists and, as a result, European alchemy 
quickly became associated with wizardry, magic, and the search for the "philosopher's stone." It wasn't 
until the late 17' A century that European chemists began applying the scientific method. Robert Boyle 
(1627 - 1691) was the first European to do so, using quantitative experiments to measure the relationship 
between the pressure and the volume of a gas. His use of the scientific method paved the way for other 
European scientists and helped to establish the modern science of chemistry. 




Figure 1.11: Antoine Lavoisier, "The Father of Modern Chemistry." 

About 100 years after Robert Boyle first performed his experiments, a French scientist by the name of 
Antoine Lavoisier (1743 - 1794) employed the scientific method when he carefully measured the masses 
of reactants and products before and after chemical reactions (Figure 1.11). Since the total mass (or 
quantity of material) never changed, Lavoisier's experiments led him to the conclusion that mass is neither 
created nor destroyed. This is known as the Law of Conservation of Mass. Lavoisier is often called 
"The Father of Modern Chemistry" because of his important contribution to the study of matter. 

After the success of Lavoisier's work, experiments involving careful measurement and observation became 
increasingly popular, leading to a rapid improvement in our understanding of chemicals and chemical 
changes. In fact, by the end of the 19 century, chemical knowledge had increased so much that practically 
everyone had stopped searching for the "philosopher's stone." 

What Chemists Do 

You might wonder why the study of chemistry is so important if you can't use it to turn iron into gold or 
to develop a potion that will make you immortal. Why didn't chemistry die when scientists like Boyle and 
Lavoisier proved alchemy was nothing but a hoax? Well, even though we can't use chemistry to make gold 



www.ckl2.org 



32 



or to live forever, modern chemistry is still very powerful! There may be no such thing as a potion that 
cures all diseases, but many chemists today are developing cures for specific diseases. In fact, chemists are 
working on everything from treatments for HIV/AIDS to medications for fighting cancer. 

Modern chemists study not only chemicals that can help us, but also chemicals that can hurt us. For 
example, environmental chemists test the air, soil, and water in our neighborhoods to make sure that 
we aren't exposed to heavy metals (such as mercury or lead) or chemical pesticides. Moreover, when 
environmental chemists do find dangerous substances, they use their knowledge of chemistry to clean up 
the contamination. Similarly, every time you buy packaged food from the grocery store, you can be sure 
that many tests have been done by chemists to make sure that those foods don't contain any toxins or 
carcinogens (cancer-causing chemicals). 

Chemists are also responsible for creating many important materials we use today. Other technologies 
rely on chemistry as well. In fact, your flat-screen LCD TV, the cubic zirconium ring on your finger, and 
the energy efficient LED lights in your home are all thanks to our improved understanding of chemistry 
(Figure 1.12). 




Figure 1.12: Energy efficient LED lights can be used to brighten your home for a party or a holiday. 

So, how do chemists accomplish all of these remarkable achievements? Unlike many of the early alchemists 
that experimented by randomly mixing together anything that they could find, today's chemists use the 
scientific method. This means that chemists rely on both careful observation and well-known physical laws. 
By putting observations and laws together, chemists develop what they term models. Models are really 
just ways of predicting what will happen given a certain set of circumstances. Sometimes these models are 
mathematical, but other times, they are purely descriptive. 

A model is any simulation, substitute, or stand-in for what you are actually studying. A good model con- 
tains the essential variables that you are concerned with in the real system, explains all the observations 
on the real system, and is as simple as possible. A model may be as uncomplicated as a sphere represent- 
ing the earth or billiard balls representing gaseous molecules, or as complex as mathematical equations 
representing light. 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. After you register the 
first time, you can return to the website (from the same computer) and view videos without registering 
again. The website has a video that applies to this lesson. The video is called "Modeling The Unseen". 
Video on Demand - Modeling the Unseen (http : //www. learner .org/resources/series61 .html?pop= 
yes&pid=793#) 

Over time, scientists have used many different models to represent atoms. As our knowledge of the atom 
changed, so did the models we use for them. Our model of the atom has progressed from an "indestructible 
sphere," to a dish of "plum pudding," to a "nuclear model." 

33 www.ckl2.org 



Hypotheses and theories comprise some ideas that scientists have about how nature works but of which 
they are not completely sure. These hypothesis and theories are models of nature used for explaining and 
testing scientific ideas. 

Chemists make up models about what happens when different chemicals are mixed together, or heated up, 
or cooled down, or compressed. Chemists invent these models using many observations from experiments 
in the past, and they use these models to predict what might happen during experiments in the future. 
Once chemists have models that predict the outcome of experiments reasonably well, those working models 
can help to tell them what they need to do to achieve a certain desired result. That result might be the 
production of an especially strong plastic, or it might be the detection of a toxin when it's present in your 
food. 

Science is not the only profession whose members make use of the scientific method. The process of making 
observations, suggesting hypotheses, and testing the hypotheses by experiment is also a common procedure 
for detectives and physicians. 

Lesson Summary 

• The word "chemistry" comes from the Arabic word "al-kimia" meaning "the art of transformation." 

• Chemistry began as the study of alchemy. Most alchemists were searching for the "philosopher's 
stone," a fabled substance that could turn common metals into gold. 

• Chinese alchemists were particularly interested in finding "the elixir of life." 

• In India, much early chemistry focused on metals. 

• The scientific method involves making careful observations and measurements and then using these 
measurements to propose hypotheses (ideas) that can, in turn, be tested with more experiments. 

• Robert Boyle and Antoine Lavoisier employed the "scientific method," thereby bringing about the 
rise of modern chemistry. 

• The Law of Conservation of Mass states that mass is neither created nor destroyed. 

• Modern chemists perform experiments and use their observations to develop models. Models then 
help chemists to understand and predict the results of future experiments. Models also help chemists 
to design new materials and cures for diseases. 

Review Questions 

1. Where does the word "chemistry" come from? 

2. Consider the following data about John's study habits, and grades: 

(a) Propose a qualitative (words, but no math) model that might describe how the length of time 
John spends studying relates to how well he does on the test? 

(b) If John wants to earn 92% on his next test, should he study for about 6 hours, 9 hours, 12 hours, 
or 18 hours? Justify your answer. 

(c) If John studies for 7 hours, do you think he will score 15%, 97%, 68%, or 48%? Justify your 
answer. 

Table 1.5: 

Hours spent studying Grade earned on the test 

20% 

5 40% 

10 60% 

www.ckl2.org 34 



Table 1.5: (continued) 



Hours spent studying Grade earned on the test 



15 



3. Helen wanted to know if lemon juice chemically reacts with tea to lighten its color. So Helen added 
25 drops of lemon juice to 250 mL of tea and observed that the tea colored lightened significantly. 
Helen wanted to make sure that the color lightening was the result of a chemical reaction and not the 
result of dilution. Which one of the following activities should Helen carry out to serve as a control 
for this experiment? 

(a) Helen should add 25 drops of orange juice to another 250 mL sample of tea. 

(b) Helen should add 25 drops of distilled water to another 250 mL sample of tea. 

(c) Helen should add 25 drops of lemon juice to a 250 mL sample of distilled water. 

(d) Helen should add 25 drops of tea to a 250 mL sample of lemon juice. 

(e) Helen should add 25 drops of tea to a 250 mL sample of tea. 

Vocabulary 

hypothesis A proposal intended to explain a set of observations. 

theory A hypothesis that has been supported with repeated testing. 

law A relationship that exists between specific observations. 

scientific method A method of investigation involving observation to generate and test hypotheses and 
theories. 

chemistry The science of the composition, structure, properties, and reactions of matter. 

1.3 Chemistry is a Science of Materials 

Lesson Objectives 

• Give examples of chemical properties a scientist might measure or observe in a laboratory. 

• Explain the difference between a physical change and a chemical change, giving examples of each. 

• Identify the situations in which mass can be converted to energy and energy can be converted to 
mass. 

Introduction 

In the last chapter we discussed some of the goals of early alchemists and some of the roles of chemists 
today. What you might have noticed is that while methods of chemical experimentation have improved 
and while knowledge of chemical properties has increased, chemistry in the 21 if century AD and chemistry 
in the 5 th century BC were both concerned with the question: How does matter change from one form to 
another? Can we predict the properties of matter? And how can we control these properties in order to 
use them to our advantage? Chemistry is essentially concerned with the science of matter and materials. 
Therefore, we'll begin our discussion of chemistry by considering some of the chemical materials that have 
been important both to early civilizations and to society today. 

35 www.ckl2.org 



Ancient Materials Versus Modern Materials 



Before humans had any understanding of chemistry, they used whatever they could find in the world around 
them. One type of material that was easily accessible to early civilizations, at least in small amounts, was 
metal. Native gold, silver, and copper, and compounds of tin and iron can all be found occurring naturally 
in cliffs and caves (in fact, the discovery of natural gold in El Dorado County, California is what lead to 
the great Gold Rush of 1849) and, as a result, these metals became very important to people in early times 
(Figure ??). 




Many ancient civilizations fashioned tools, jewelry, and weapons out of metal that they scavenged from rocks 
around them (Figure ??). After a while, however, people discovered that by mixing naturally occurring 
metals with other substances, they could create new materials that often had superior properties. 




Some of the oldest materials produced by man include mixtures (or more specifically solutions) of metals 
known as alloys. One of the earliest alloys ever discovered was bronze. Bronze can be made by heating 
chunks of tin and copper until they are liquid and then mixing the two pure metals together. Bronze was 
very important to early civilizations because it was more resistant to rust than iron, harder than copper, 
and could hold an edge and be sharpened to create tools and weapons. 

Another alloy, produced early in the history of civilization, is steel. As you learned in an earlier section, 
steel is an alloy of iron and carbon (or charcoal). Steel, particularly Wootz steel (which required a special 
technique that involved the addition of glass), was especially strong, and could be fashioned into very 
sharp edges, perfect for swords. Another old material whose production was known to early civilizations 
is brass. Again, brass is an alloy, made of two pure metals, copper, and zinc. Early Romans knew that if 
they melted copper, and a zinc ore known as calamine together, they could produce brass, which was both 



www.ckl2.org 



36 



shiny like gold, and resistant to rust. Brass was a common material used to make coins. 

What you might notice about these "old" materials is that they are mainly alloys. At the time when bronze 
and brass and steel were discovered, people didn't know much about the composition of matter or about 
how matter was assembled on a microscopic scale. As a result, inventing materials was largely a matter of 
trial and error. Towards the end of the 19 century, however, scientists were beginning to understand the 
make-up of matter, and this understanding led to new insight into how to develop materials with desirable 
properties. 




8 JL* 



One of the huge breakthroughs in recent history has been the discovery of plastic and plastic products 
(Figure ??). Initially, plastic was made by chemically modifying cellulose, a naturally occurring chemical 
found in plants. As chemical knowledge developed, however, scientists began to realize that plastics had 
special properties because, on a microscopic scale, they were composed of thousands of tiny chains of 
molecules all tangled up together. Scientists reasoned that if they altered the chemicals in these chains, 
but still managed to keep the chains intact, they could make new plastics with new properties. Thus began 
the plastic revolution! 

Semiconductors are another class of "new" materials whose development is largely based on our improved 
understanding of chemistry. Because scientists know how matter is put together, they can predict how to 
fine-tune the chemical composition of a semiconductor in order to make it absorb light and act as a solar 
cell or emit light and act as a light source. We've come a long way from our early days of producing bronze 
and steel. Nevertheless, as our understanding of chemistry improves, we will be able to create even more 
useful materials than we have today. 

Chemists Study the Properties of Matter 

Hopefully at this point you are fully convinced of how important and useful the study of chemistry can 
be. You may, however, still be wondering exactly what it is that a chemist does. Chemistry is the study of 
matter and the changes that matter undergoes. In general, chemists are interested in both characteristics 
that you can test and observe, like a chemical's smell or color, and characteristics that are far too small to 
see, like what the oxygen you breathe in or the carbon dioxide you breath out looks like under a microscope 
1,000 times more powerful than any existing in the world today. 




Wait a minute... how can a chemist know what oxygen and carbon dioxide look like under a microscope 
that doesn't even exist? What happened to the scientific method? What happened to relying on observa- 
tions and careful measurements? In fact, because chemists can't see the underlying structure of different 

37 www.ckl2.org 



materials, they have to rely on the scientific method even more! Chemists are a lot like detectives (Figure 
??). Suppose a detective is trying to solve a murder case - what does she do? Obviously, the detective 
starts by visiting the site of the crime and looking for evidence. If the murderer has left enough clues 
behind, the detective can piece together a theory explaining what happened. 

Even though the detective wasn't at the crime scene when the crime was committed and even though 
the detective didn't actually see the murderer kill the victim, with the right evidence, the detective can 
be pretty sure he or she knows how it took place. It's the same with chemistry. When chemists go into 
the laboratory, they collect evidence by making measurements. Once they've collected enough clues from 
the properties that they can observe, they use that evidence to piece together a theory explaining the 
properties that they can't observe - the properties that are too small to see. 

What kinds of properties do chemists actually measure in the laboratory? Well, you can probably guess a 
few. Imagine that you go to dinner at a friend's house and are served something that you don't recognize - 
what types of observations might you make to determine exactly what you've been given? You might smell 
the food. You might note the color of the food. You might try to decide whether the food is a liquid or a 
solid because if it's a liquid, it's probably soup or a drink. The temperature of the food could be useful if 
you wanted to know whether or not you'd been served ice cream! You could also pick up a small amount 
of food with your fork and try to figure out how much it weighs - a light dessert might be something like 
an angel cake, while a heavy dessert is probably a pound cake. The quantity of food you've been given 
might be a clue too. Finally, you might want to know something about the food's texture - is it hard and 
granular like sugar cubes, or soft and easy to spread, like butter? 

Believe it or not, the observations you are likely to make when trying to identify an unknown food are very 
similar to the observations that a chemists makes when trying to learn about a new material. Chemists 
rely on smell, color, state (that is, whether it is a solid, liquid or gas), temperature, volume, mass (which 
is related to weight, as you'll discover in a later section), and texture. There is, however, one property you 
might use to learn about a food, but that you should definitely not use to learn about a chemical - taste! 

In The Atomic Theory, you'll see exactly how measurements of certain properties helped early scientists to 
develop theories about the chemical structure of matter on a scale much smaller than they could ever hope 
to see. You'll also learn how these theories, in turn, allow us to make predictions about new materials that 
we haven't even created yet. 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. After you register the 
first time, you can return to the website (from the same computer) and view videos without registering 
again. The website has two videos that apply to this lesson. One is a video called The World of 
Chemistry that relates chemistry to other sciences and daily life. Another video called Thinking Like 
Scientists relates to the scientific method. The audience on the video is young children but the ideas 
are full grown. Video on Demand - The World of Chemistry (http://www.learner.org/resources/ 
series61.html?pop=yes&pid=793#) 



Chemists Study of How and Why Matter Changes 

In the last section, we discussed the properties of matter and how scientists use these properties to deduce 
certain facts about the structure of matter. However, if chemists only studied properties such as color and 
smell, they would only be collecting half of the evidence. While the properties of matter can tell us a lot, 
so too can the changes that matter undergoes. Suppose you've been served a slice of cake that you've 
noticed is cold to the touch (Figure ??). You might guess that you're dealing with an ice cream cake. 
But then again, maybe it's just a normal cake that's been kept in the freezer. Can you think of some way 
to tell between a frozen slice of ice cream cake and a frozen slice of regular cake? Well, one possibility is 



www.ckl2.org 38 



to wait for a while and see whether your slice melts. If the slice melts, it was an ice cream cake, and if it 
doesn't, it was just regular cake. In this case, you aren't observing a property, but rather a change in a 
property. The property being changed in the example is state. 




Similarly, chemists learn a lot about the nature of matter by studying the changes that matter can undergo. 
Chemists make a distinction between two different types of changes that they study - physical changes and 
chemical changes. Physical changes are changes that do not alter the identity of a substance. Some 
types of physical changes include: 

• Changes of state (changes from a solid to a liquid or a gas and vice versa) 

• Separation of a mixture 

• Physical deformation (cutting, denting, stretching) 

• Making solutions (special kinds of mixtures) 



When you have a jar containing a mixture of pennies and nickels and you sort the mixture so that you 
have one pile of pennies and another pile of nickels, you have not altered the identity of either the pennies 
or the nickels - you've merely separated them into two groups. This would be an example of a physical 
change. Similarly, if you have a piece of paper, you don't change it into something other than a piece of 
paper by ripping it up. What was paper before you starting tearing is still paper when you're done. Again, 
this is an example of a physical change (Figure ??). 




You might find it a little harder to understand why changes in state are physical changes. Until we discuss 
chemicals in terms of the smaller units (atoms and molecules) that make them up, it probably won't be 
clear to you why freezing a substance or boiling a substance is only a physical change. 

For now, though, you just have to trust that changes in state are physical changes. If you're ever in doubt, 
remember this: when a lake freezes in the winter, the water doesn't disappear or turn into something else 
- it just takes on a new form. Liquid water and solid water (ice) are just different forms of the substance 
we know as water. For the most part, physical changes tend to be reversible - in other words, they can 
occur in both directions. You can turn liquid water into solid water through cooling; you can also turn 
solid water into liquid water through heating. 

The other type of change that chemists are concerned with is chemical change. A chemical change occurs 
when one substance is turned into an entirely new substance as a result of a chemical reaction (Figure ??). 



39 



www.ckl2.org 



Again, as we learn more about chemicals, and what chemicals look like, the meaning of a chemical change 
and the distinction between a chemical change and a physical change will become more obvious. For now, 
realize that chemicals are made up of tiny units known as atoms. Some of these atoms are bonded (or 
"glued") together, but during a chemical change, some of the bonds are broken and new bonds are formed. 

You're probably wondering how you know when a chemical change has occurred. Sometimes it can be 
pretty tricky to tell, but there are several evidences of chemical changes to look for. There has probably 
been a chemical change if: 

• A change in color has occurred 

• Light, heat or sound has been given off from the material itself 

• A precipitate (a solid formed when two liquids are mixed) has appeared 

• A gas has been produced (detected by bubbling or a new odor) 




Chemical changes are frequently harder to reverse than physical changes. One good example of a chemical 
change is burning paper. In contrast to the act of ripping paper, the act of burning paper actually results in 
the formation of new chemicals (carbon dioxide and water, to be exact). Notice that whereas ripped paper 
can be at least partially reassembled, burned paper cannot be "unburned." In other words, burning only 
goes in one direction. The fact that burning is not reversible is another good indication that it involves a 
chemical change. 

Chemists Study the Interchange of Matter and Energy 

Chemists are concerned with the properties of matter, and the changes that matter undergoes. For the 
most part, though, the changes that chemists are interested in are either physical changes, like changes in 
state, or chemical changes like chemical reactions. In either case, Lavoisier's Law of Conservation of Mass, 
applies. In both physical and chemical changes, matter is neither created nor destroyed. There is, however, 
another type of change that matter can undergo that actually disobeys Lavoisier's Law of Conservation of 
Mass, and that is the conversion of matter into energy, and vice versa. 

Back when Lavoisier was studying chemistry, the technology didn't exist that would allow scientists to turn 
matter into energy and energy into matter, but it can be done. This is the concept that Einstein proposed 
in his famous equation E = mc 2 . (This equation states that the energy in a given amount of matter is equal 
to the mass of the matter times the speed of light squared.) Chemical reactions don't involve changing 
measureable amounts of energy to mass or mass to energy. Mass-energy conversion is, however, important 
in chemistry that deals with radioactivity and particularly in the production of electricity by nuclear power 
plants. 

Lesson Summary 

• Some of the earliest materials invented by humans were alloys such as bronze, steel, and brass. 

• With improved understanding of chemistry comes the ability to design new and useful materials, like 
plastics and semiconductors. 

www.ckl2.org 40 



• Chemists can't actually see the underlying structure of most materials. As a result, they measure 
properties that they can see or observe and use this evidence to develop theories that explain how 
chemicals are organized on a sub-microscopic (smaller than you can see with a microscope) scale. 

• Some of the physical properties that scientists observe pertain to state (solid, liquid, or gas), tem- 
perature, volume, mass, and texture. 

• Chemists also study the changes that different materials undergo; this can give them valuable infor- 
mation about the chemicals involved. 

• There are two types of changes that are important in chemistry - physical changes and chemical 
changes. 

• Physical changes are changes that do not alter the identity of a substance; they are usually reversible. 

• Chemical changes are changes that occur when one substance is turned into another substance as a 
result of a chemical reaction. They are usually difficult to reverse. 

• It is also possible to change matter into energy and energy into matter. 

Review Questions 

1. Name the two types of changes that chemists are primarily interested in. 

2. Decide whether each of the following statements is true or false. 

(a) Plastics were developed in Rome around 300 AD 

(b) Bronze is an example of an alloy 

(c) Plastic is an example of an alloy 

(d) Brass is an example of an alloy 

3. Match the following alloys with their common names. 

(a) Bronze - a. tin and copper 

(b) Brass - b. is not an alloy 

(c) Plastic - c. iron and carbon 

(d) Tin - d. copper and zinc 

(e) Steel - e. is not an alloy 

4. Decide whether each of the following statements is true or false. 

(a) Physical changes are typically accompanied by a color change 

(b) A burning campfire is an example of a chemical change 

(c) When you heat your house with coal, the coal undergoes a chemical change 

(d) When you drop a plate, and it breaks, the plate undergoes a physical change 

5. In each of the following examples, determine whether the change involved is a physical change or a 
chemical change. 

(a) Flattening a ball of silly putty 

(b) Combining a bowl of cherries and a bowl of blueberries 

(c) Boiling water 

(d) Cooking an egg 

6. Judy has two beakers filled with clear liquids, and she needs to know whether the liquid in the 
first beaker is the same as the liquid in the second beaker. In which scenario does Judy use physical 
properties to answer her question, and in which scenario does Judy use changes in chemical properties 
to answer her question? 

(a) Judy smells the two liquids and notices that the liquid in the first beaker has a strong odor, 
while she can't smell the liquid in the second beaker at all. 

(b) Judy mixes some table salt into the first beaker and notices that a white precipitate forms. She 
then mixes some table salt into the second beaker, but nothing happens. 

41 www.ckl2.org 



Vocabulary 

alloy A solution (or a special kind of mixture), in which at least one of the components is a metal. 

physical change Changes that do not alter the identity of the substance. 

chemical change A change that occurs when one substance is turned into an entirely new substance as 
a result of a chemical reaction. 

1.4 Matter 

Lesson Objectives 

• Define matter and explain how it is composed of building blocks known as "atoms." 

• Distinguish between mass and weight. 

Introduction 

We are all familiar with matter. The definition of matter is anything that has mass and volume (takes 
up space). For most common objects that we deal with every day, it is fairly simple to demonstrate that 
they have mass and take up space. You might be able to imagine, however, the difficulty for people several 
hundred years ago to demonstrate that air has mass and volume. Air (and all other gases) are invisible 
to the eye, have very small masses compared to equal amounts of solids and liquids, and are quite easy to 
compress (change volume). Without sensitive equipment, it would have been difficult to convince people 
that gases are matter. Today, we can measure the mass of a small balloon when it is deflated and then blow 
it up, tie it off, and measure its mass again to detect the additional mass due to the air inside. The mass 
of air, under room conditions, that occupies a one quart jar is approximately 0.0002 pounds. This small 
amount of mass would have been difficult to measure in times before balances were designed to accurately 
measure very small masses. Later, scientists were able to compress gases into such a small volume that 
the gases turned into liquids, which made it clear that gases are matter. 

On the other hand, when you add heat to an object, the temperature of the object increases, but even the 
most sensitive balance cannot detect any difference in mass between an object when cold and when hot. 
Heat does not qualify as matter. 

The Material in the Universe 

Knowing that planets, solar systems, and even galaxies are made out of matter doesn't bring us any closer 
to understanding what matter is. Up until the early 1800s, people didn't really understand matter at all. 
They knew that there were "things" in the world that they could pick up and use, and that some of these 
"things" could be turned into other "things." For example, someone who found a piece of copper could 
shape it into a necklace or melt it together with zinc to make brass. What people didn't know, though, 
was how all of these "things" were related. If they had, the alchemists probably wouldn't have wasted so 
much time trying to convert common metals into gold. You'll understand why by the time you're finished 
with this section. Even though the universe consists of "things" as wildly different as ants and galaxies, 
the matter that makes up all of these "things" is composed of a very limited number of building blocks 
(Figure 1.13). 

www.ckl2.org 42 




Figure 1.13: Everything from an ant to an entire galaxy is composed of matter. 



These building blocks are known as atoms, and so far, scientists have discovered or created a grand total 
of 117 different types of atoms. Scientists have given a name to each different type of atom. A substance 
that is composed of only one type of atom is called an element. Each element, therefore, has its own name; 
it also has its own symbol. The "periodic table" is a way of summarizing all of the different atoms that 
scientists have discovered (Figure 1.14). Each square in the periodic table contains the symbol (a capital 
letter or a capital letter followed by a lower case letter) for one of the elements. 



IA 
1 



alkali metals 



H 



post-transition metals [ 



alkaline earths 



transition metals 



1MB IVB VB VIB /MB VIII VIII 
3 4 5 6 W 8 9 



VIII IB\ MB 
10 11 \l2 





noble aases 




IIIA 
13 


IVA VA VIA 
14 15 16 


VIIA 
17 | 


B 


C N O 


F 


Al 


Si P S 


C, 



VIIA 



Li Be 
Na Mg 

K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br 
Cs Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn,Sb Te I 



Rb Ba La 
Fr Ra Ac 



Hf Ta W Re Os Ir Pt Au 



Rf 



Db 



Sg 



Bh Hs Mt Ds 



Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu 




lanthanides 



Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr 



antinides I 



S.K. Lower 



Figure 1.14: The Periodic Table. 

At this point, what should amaze you is that all forms of matter in our universe are made with only 
117 different building blocks. In some ways, it's sort of like cooking a gourmet, five-course meal using 
only three ingredients! How is it possible? To answer that question, you have to understand the ways in 
which different elements are put together to form matter. The most important method that nature uses 
to organize atoms into matter is the formation of molecules. Molecules are groups of two or more atoms 
that have been bonded together. There are millions of different ways to bond atoms together, which means 
that there are millions of different possible molecules. Each of these molecules has its own set of chemical 
properties, and it's these properties with which chemists are most concerned. You will learn a lot more 
about atoms and molecules, including how they were discovered, in a later part of the textbook. Figure 
1.15, however, gives you a preview of some of the common molecules that you might come in contact with 
on a daily basis. 

Now back to the question of why alchemists had trouble making gold out of other common metals. Most 
naturally occurring metals, including gold, iron, copper, and silver are elements. Look carefully at the 
Periodic Table in Figure 1.14. Do you see the symbol Au? Au is the symbol for gold. Gold is one of the 



43 



www.ckl2.org 



oxygen 

k 4 

hydrogen 
a water molecule 



oxygen 




oxygen 



CM) 

carbon 
a carbon dioxide molecule 



a nitrogen dioxide molecule 



nitrogen 



hydrogen 
an ammonia molecule 



oxygen 




an oxjgen molecule 



Figure 1.15: Some common molecules and the atoms that they are made from. 



www.ckl2.org 



44 



117 different types of atoms - it's one of nature's building blocks. Take another look at the Periodic Table. 
Do you see the symbol Fe? Fe is the symbol for iron. Again, that means that iron is an element, or a type 
of atom. In fact, copper (Cm), tin (Sn), and silver (Ag) are all elements. In other words, alchemists were 
trying to convert one type of element, or building block, into another type of element, and that just can't 
be done by chemical means. 

Chemical reactions can turn elements into molecules, molecules into other molecules, and molecules back 
into elements. Chemical reactions cannot, however, turn one type of element into another type of element. 
The only way to do that is through what are known as nuclear reactions, and nuclear reactions require 
advanced technical equipment that wasn't around in the days of the alchemists. It's like building a house. 
You can make a house by cementing together bricks, stones, and wood, just like you can make a molecule 
by bonding together different types of atoms. You can also get your bricks, stones, and wood back by 
taking the house apart just like you can get your atoms back by taking the molecule apart. No matter how 
hard you try, though, you can't turn bricks into wood. Converting common metals like copper into gold 
or iron into gold would be like turning bricks into wood - it's simply not possible. 



Matter Has Mass and Occupies Space 

So far we've decided that the entire universe is composed of matter, which is in turn composed of atoms. 
Frequently, though, chemists want to know how much matter they actually have. To figure this out they 
rely on two fundamental properties of matter. All matter in the universe, from a teaspoon of salt to the 
Pacific Ocean, has mass and occupies space. When scientists measure how much space is taken up by a 
certain quantity of matter, they are measuring the objects volume. Obviously, the volume of the Pacific 
Ocean is a lot larger than the volume of a teaspoon of salt. Unfortunately, while volume is an important 
property, and plays an important role in a lot of different chemical experiments, volume is not the best 
way to determine how much matter you have. 

Typically we think that the bigger something is, the more there is in it. That's certainly true a lot of 
the time in our everyday lives. If you pour yourself two cups of coffee in the morning, you'll be drinking 
twice as much coffee as you would have if you'd only poured yourself a single cup. Unfortunately, any time 
that we compare volumes in this way, we are making two assumptions that aren't always true in chemical 
experiments. First, we are assuming constant temperature. That's important, because the amount of space 
taken up by a certain quantity of matter depends on the temperature of that matter. In general, heating 
something up causes it to expand, and cooling something down causes it to contract. Secondly, when you 
compare volumes in everyday life, you are almost always comparing volumes of the same material. You 
can compare two cups of coffee to one cup of coffee, but how do you compare two cups of coffee to one cup 
of ice cream? It really doesn't make sense. Volume is not a good way to determine the quantity of matter 
that you have. 

If you can't use volume to figure out how much matter you've got, what can you use? It turns out that 
the best way to determine quantities of matter is to use a measure known as mass. The mass of an object 
doesn't change with temperature, which makes it a lot easier to determine how much stuff you're dealing 
with, especially when you don't know what the temperature is or when the temperature keeps changing. 
Another good thing about mass is that an atom of a particular element always has the same mass (Strictly 
speaking that's not entirely true because of what are known as isotopes, but you won't need to worry about 
that now). For example, an atom of gold always has a mass of about 197 atomic mass units or daltons. 
Atomic mass units are units that we use to measure mass, just like a mile is a unit that we use to measure 
distance, and an hour is a unit that we use to measure time. Even when atoms are bonded together into 
molecules, the individual atoms have the same mass, meaning that by adding up all of the masses of the 
atoms in a molecule, it's fairly easy to figure out the mass of the molecule itself. You'll eventually learn 
how to do this. 



45 www.ckl2.org 



The Difference Between Mass and Weight 

One typical mistake that students make when learning about mass for the first time is confusing mass with 
weight. Again, this confusion is largely due to the fact that, in everyday English, we frequently use the 
word "weight" when we actually mean "mass." For example, when you say "I want to lose weight," what 
you really mean is, "I want to lose mass." In science, the word "weight" has a very specific definition that 
is different from what you might expect. Do not confuse the everyday meaning of the word "weight" with 
the scientific meanings of the word "weight" (Figure 1.16). 




Figure 1.16: The type of balance a chemist might use to measure the mass of an object. 

In science, mass is an intrinsic ("built-in") property of matter. The mass of an atom is the same regardless 
of the temperature or the other atoms that are bonded to it. 




Figure 1.17: With no large mass nearby, this spaceship is weightless. 

Similarly, the mass of an atom doesn't change depending on where it is. The mass of an atom is the same 
on Earth as it would be on the moon, or on Jupiter, or in the middle of space. Weight, on the other hand, 
does change with location. An object that weighs 240 pounds on earth would weigh about 40 pounds on 
the moon and that same object in a spaceship far away from any large mass would weigh zero (Figure 
1.17). In all three cases, however, the object would have exactly the same mass. 

In science, weight is a measurement of how strongly gravity pulls on an object. Weight depends on both 
the mass of the object and the force of gravity the object is experiencing. That's why your weight changes 
depending on where you're standing. In each case, your mass will be the same, but the force of gravity 
on Earth is different than the force of gravity on the moon, or on Jupiter and, as a result, your weight is 
different too. The force of gravity, however, doesn't change significantly on the surface of Earth. In other 



www.ckl2.org 



46 



words, the force of gravity in California is approximately the same as the force of gravity in Australia. As 
a result, as long as you stick to the surface of the Earth, the more massive an object is, the more it weighs. 

Lesson Summary 

• All physical objects are made of matter. 

• Matter itself is composed of tiny building blocks known as "atoms." There are only 117 different types 
of atoms known to man. 

• Frequently, atoms are bonded together to form "molecules." 

• All matter has mass and occupies space. 

• Volume is a measure of how much space an object occupies. Volume is not a good measure of how 
much matter makes up any given object. 

• Mass is an intrinsic property of matter that does not depend on temperature, location, or the way in 
which the matter is organized (how the atoms are bonded) As a result, mass is an excellent measure 
of how much matter is in any given object. 

• "Mass" and "weight" have two very different scientific meanings. 

• "Mass" only depends on how much matter is in an object. "Weight," on the other hand, depends on 
how strongly gravity pulls on an object. 

Review Questions 

1. What is matter? 

2. In this chapter, we'll learn about atoms, which are the building blocks of all matter in the universe. 
As of 2007, scientists only know of 117 different types of atoms. How do you think it's possible to 
generate so many different forms of matter using only 117 types of building blocks? 

3. Which do you think has more matter, a cup of water or a cup of mercury? Explain. 

4. Decide whether each of the following statements is true or false. 

(a) Mass and weight are two words for the same concept. 

(b) Molecules are bonded together to form atoms. 

(c) Alchemists couldn't make gold out of common metals because gold is an element. 

(d) The symbol for Gold in the periodic table is Gd. 

5. Would you have more mass on the moon or on Earth? 

6. Would you have more weight on the moon or on Earth? The force of gravity is stronger on the Earth 
than it is on the moon. 

7. Match the following terms with their meaning. 

(a) Mass - a. a measure of the total quantity of matter in an object 

(b) Volume - b. a measure of how strongly gravity pulls on an object 

(c) Weight - c. a measure of the space occupied by an object 

8. For the following statements, circle all of the options that apply: Mass depends on... 

(a) the total quantity of matter 

(b) the temperature 

(c) the location 

(d) the force of gravity 

Volume depends on... 

(a) the total quantity of matter 

(b) the temperature 

47 www.ckl2.org 



(c) the object's shape (independent of size) 

(d) the object's size (independent of shape) 

Weight depends on... 

(a) the total quantity of matter 

(b) the temperature 

(c) the location 

(d) the force of gravity 

Further Reading / Supplemental Links 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 1-4 is on the Classification of Matter. 

• Lesson 1-5 is on Physical and Chemical Properties and Changes, http://www.fordhamprep.org/ 
gcurran/sho/sho/lessons/lessonl4 . htm 

Vocabulary 

matter Anything of substance that has mass and occupies space. 

atom The basic building block of all matter. There are 117 known types of atoms. While atoms can be 
broken down into particles known as electrons, protons and neutrons, this is very difficult to do. 

element A type of atom. There are 117 known elements. 

molecule Two or more atoms bonded together. Specific molecules, like water, have distinct characteris- 
tics. 

Periodic Table A way of summarizing all the different atoms that scientists have discovered. Each 
square in the periodic table contains the symbol for one of the elements. 

mass An intrinsic property of matter that can be used to measure the quantity of matter present in a 
sample. 

weight A measurement of how strongly gravity pulls on an object. 

volume A measurement of how much space a substance occupies. 

1.5 Energy 

Lesson Objectives 

• Define heat and work. 

• Distinguish between kinetic energy and potential energy. 

• State the law of conservation of matter and energy. 

www.ckl2.org 48 



Introduction 

Just like matter, energy is something that we are all familiar with and use on a daily basis. Before you 
go on a long hike, you eat an energy bar; every month, the energy bill is paid; on TV, politicians argue 
about the energy crisis. But have you ever wondered what energy really is? When you plug a lamp into an 
electric socket, you see energy in the form of light, but when you plug a heating pad into that same socket, 
you only feel warmth. When you eat a bowl of spaghetti, the energy it provides helps you to function 
throughout the day, but when you eat five bowls of spaghetti, some of that energy is turned into body fat. 

If you stop to think about it, energy is very complicated. Still, we use energy for every single thing that 
we do, from the moment we wake up to the moment we go to sleep. Without energy, we couldn't turn on 
lights, we couldn't brush our teeth, we couldn't make our lunch, and we couldn't travel to school. Although 
we all use energy, very few of us understand how we use it. 

Ability to Do Work or Produce Heat 

When we speak of using energy, we are really referring to transferring energy from one place to another. 
When you use energy to throw a ball, you transfer energy from your body to the ball, and this causes 
the ball to fly through the air. When you use energy to warm your house, you transfer energy from the 
furnace to the air in your home, and this causes the temperature in your house to rise. Although energy is 
used in many kinds of different situations, all of these uses rely on energy being transferred in one of two 
ways. Energy can be transferred as heat or as work. Unfortunately, both "heat" and "work" are common 
words, so you might think that you already know their meanings. In science, the words "heat" and "work" 
have very specific definitions that are different from what you might expect. Do not confuse the everyday 
meanings of the words "heat" and "work" with the scientific meanings. 

When scientists speak of heat, they are referring to energy that is transferred from an object with a higher 
temperature to an object with a lower temperature as a result of the temperature difference. Heat will 
"flow" from the hot object to the cold object until both end up at the same temperature. When you cook 
with a metal pot, you witness energy being transferred in the form of heat. Initially, only the stove element 
is hot - the pot and the food inside the pot are cold. As a result, heat moves from the hot stove element to 
the cold pot. After a while, enough heat is transferred from the stove to the pot, raising the temperature 
of the pot and all of its contents (Figure 1.18). 




Figure 1.18: Energy is transferred as heat from the hot stove element to the cooler pot until the pot and 
its contents become just as hot as the element. The energy that is transferred into the pot as heat is then 
used to cook the food. 

49 www.ckl2.org 



We've all observed heat moving from a hot object to a cold object, but you might wonder how the energy 
actually travels. Whenever an object is hot, the molecules within the object are shaking and vibrating 
vigorously. The hotter an object is, the more the molecules jiggle around. As you'll learn in the next 
section, anything that is moving has energy, and the more it's moving, the more energy it has. Hot objects 
have a lot of energy, and it's this energy that is transferred to the colder objects when the two come in 
contact. The easiest way to visualize heat transfer is to imagine a domino effect. 

Heat is being transferred from a hot object to a colder object. In detail: a. As the red molecules in the hot 
object jiggle and vibrate, they hit some of the blue molecules in the colder object. This transfers energy 
from the hot molecules to the colder molecules, causing these molecules to vibrate faster, b. - d. Just 
like dominoes, heat passes along the chain until the energy is spread equally between all of the molecules. 
(Source: Sharon Bewick. CC-BY-SA) 



hot object 



cold object 



cfp- 



■o o o 



y * * 






o o 



it ~~% 

o-£H 



<t 



ti 



o*;yo 









Take a close look at the figure above. When the vibrating molecules of the hot object bump into the 
molecules of the colder object, they transfer some of their energy, causing the molecules in the colder 
object to start vibrating vigorously as well. As these molecules vibrate, they bump into their neighbors 
and transfer some of their energy on down the chain. In this way, energy passes through the whole system 
until all of the molecules have about the same amount, and the initial objects are at the same temperature. 

Heat is only one way in which energy can be transferred. Energy can also be transferred as work. The 
scientific definition of work is force (any push or pull) applied over a distance. Whenever you push an object 
and cause it to move, you've done work, and you've transferred some of your energy to the object. At this 
point, it's important to warn you of a common misconception. Sometimes we think that the amount of 
work done can be measured by the amount of effort put in. This may be true in everyday life, but it isn't 
true in science. By definition, scientific work requires that force be applied over a distance. It doesn't 
matter how hard you push or how hard you pull. If you haven't moved the object, you haven't done any 
work. 

So far, we've talked about the two ways in which energy can be transferred from one place, or object, to 
another. Energy can be transferred as heat, and energy can be transferred as work. But the question still 
remains - what IS energy? We'll try to at least partially tackle that question in the next section. 



www.ckl2.org 



50 



Types of Energy: Kinetic and Potential 

Machines use energy, our bodies use energy, energy comes from the sun, energy comes from volcanoes, 
energy causes forest fires, and energy helps us to grow food. With all these seemingly different types of 
energy, it's hard to believe that there are really only two different forms of energy - kinetic energy and 
potential energy. Kinetic energy is energy associated with motion. When an object is moving, it has 
kinetic energy. When the object stops moving, it has no kinetic energy. While all moving objects have 
kinetic energy, not all moving objects have the same amount of kinetic energy. The amount of kinetic 
energy possessed by an object is determined by its mass, and its speed. The heavier an object is and the 
faster it is moving, the more kinetic energy it has. 

Kinetic energy is very common, and it's easy to spot examples of it in the world around you. Sometimes 
we even try to capture kinetic energy and use it to power things like our home appliances. If you're from 
California, you might have driven through the Tehachapi Pass near Mojave or the Montezuma Hills in 
Solano County and seen the windmills lining the slopes of the mountains (Figure 1.19). These are two of 
the larger wind farms in North America. As wind rushes along the hills, the kinetic energy of the blowing 
air particles turns the windmills, trapping the wind's kinetic energy so that people can use it in their houses 
and offices. 




Figure 1.19: A wind farm in the Tehachapi Mountains of Southern California. Kinetic energy from the 
rushing air particles turns the windmills, allowing us to capture the wind's kinetic energy and use it. 

Capturing kinetic energy can be very effective, but if you think carefully, you'll realize that there's a small 
problem. Kinetic energy is only available when something is moving. When the wind is blowing, we can 
use its kinetic energy, but when the wind stops blowing, there's no kinetic energy available. Imagine what 
it would be like trying to power your television set using the wind's kinetic energy. You could turn on the 
TV and watch your favorite program on a windy day, but every time the wind stopped blowing, your TV 
screen would flicker off because it would run out of energy. You'd probably only be able to watch about 
half of the episodes, and you'd never know what was going on! 

Of course, when you turn on the TV, or flip on the lights, you can usually count on them having a constant 
supply of energy. This is largely because we don't rely on kinetic energy alone for power. Instead, we use 
energy in its other form - we use potential energy. Potential energy is stored energy. It's energy that 
remains available until we choose to use it. Think of a battery in a flashlight. If you leave a flashlight on, 
the battery will run out of energy within a couple of hours, and your flashlight will die. If, however, you 
only use the flashlight when you need it, and you turn it off when you don't, the battery will last for days 
or even months. The battery contains a certain amount of energy, and it will power the flashlight for a 
certain amount of time, but because the battery stores potential energy, you can choose to use the energy 

51 www.ckl2.org 



all at once, or you can save it and only use a small amount at a time. 

Any stored energy is potential energy. Unfortunately, there are a lot of different ways in which energy 
can be stored, and that can make potential energy very difficult to recognize. In general, an object has 
potential energy because of its position relative to another object. For example when you hold a rock above 
the earth, it has potential energy because of its position relative to the ground. You can tell that this is 
potential energy because the energy is stored for as long as you hold the rock in the air. Once you drop 
the rock, though, the stored energy is released. 

There are other common examples of potential energy. A ball at the top of a hill stores potential energy 
until it is allowed to roll to the bottom. When you hold two magnets next to each other, they store 
potential energy too. For some examples of potential energy, though, it's harder to see how "position" is 
involved. In chemistry, we are often interested in what is called chemical potential energy. Chemical 
potential energy is energy stored in the atoms, molecules, and chemical bonds that make up matter. How 
does this depend on position? 

As you learned earlier, the world, and all of the chemicals in it are made up of atoms and molecules. These 
store potential energy that is dependent on their positions relative to one another. Of course, you can't see 
atoms and molecules. Nevertheless, scientists do know a lot about the ways in which atoms and molecules 
interact, and this allows them to figure out how much potential energy is stored in a specific quantity (like 
a cup or a gallon) of a particular chemical (Figure 1.20). Different chemicals have different amounts of 
potential energy because they are made up of different atoms, and those atoms have different positions 
relative to one another. 






Two hydrogen atoms and an 
oxygen atom making up a 
water molecule 



Figure 1.20: Scientists use their knowledge of what the atoms and molecules look like and how they interact 
to determine the potential energy that can be stored in any particular chemical substance. 

Since different chemicals have different amounts of potential energy, scientists will sometimes say potential 
energy depends not only on position, but also on composition. Composition affects potential energy 
because it determines which molecules and atoms end up next to each other. For example, the total 
potential energy in a cup of pure water is different than the total potential energy in a cup of apple juice, 
because the cup of water and the cup of apple juice are composed of different amounts of different chemicals. 

At this point, you might be wondering just how useful chemical potential energy is. If you want to release 
the potential energy stored in an object held above the ground, you just drop it. But how do you get 
potential energy out of chemicals? It's actually not that difficult. You use the fact that different chemicals 
have different amounts of potential energy. If you start with chemicals that have a lot of potential energy 
and allow them to react and form chemicals with less potential energy, all the extra energy that was in the 
chemicals at the beginning but not at the end is released. 

www.ckl2.org 52 



Law of Conservation of Matter and Energy 

So far we've talked about how energy exists as either kinetic energy or potential energy and how energy 
can be transferred as either heat or work. While it's important to understand the difference between 
kinetic energy and potential energy and the difference between heat and work, the truth is, energy is 
constantly changing. Kinetic energy is constantly being turned into potential energy, and potential energy 
is constantly being turned into kinetic energy. Likewise, energy that is transferred as work might later end 
up transferred as heat, while energy that is transferred as heat might later end up being used to do work. 

Even though energy can change form, it must still follow one fundamental law - Energy cannot be created or 
destroyed, it can only be changed from one form to another. This law is known as the Law of Conservation 
of Energy. In a lot of ways, energy is like money. You can exchange quarters for dollar bills and dollar 
bills for quarters, but no matter how often you convert between the two, you won't end up with any more 
or any less money than you started with. Similarly, you can transfer (or spend) money using cash, or 
transfer money using a credit card, but you still spend the same amount of money, and the store still 
makes the same amount of money. 

As it turns out, the law of conservation of energy isn't exactly the whole truth. If you think back, you'll 
remember that energy and matter are actually interchangeable. In other words, energy can be created 
(made out of matter) and destroyed (turned into matter). As a result, the law of conservation of energy 
has been changed into the Law of Conservation of Matter and Energy. This law states that 

The total amount of mass and energy in the universe is conserved (does not change). 

This is one of the most important laws you will ever learn. Nevertheless, in chemistry we are rarely 
concerned with converting matter to energy or energy to matter. Instead, chemists deal primarily with 
converting one form of matter into another form of matter (through chemical reactions) and converting 
one form of energy into another form of energy. 

Let's take a look at several examples, where kinetic energy is switched to potential energy and vice versa. 
Remember Wile E. Coyote with his anvil poised at the top of the cliff? As long as Wile E. Coyote holds 
the anvil and waits for Road Runner, the anvil stores potential energy. However, when Wile E. Coyote 
drops the anvil, the original potential energy stored in the anvil is converted to kinetic energy. The further 
the anvil falls, the faster it falls, and more and more of the anvil's potential energy is converted to kinetic 
energy. 

The opposite, of course, happens when you throw a ball into the air. When the ball leaves your hand, it 
has a lot of kinetic energy, but as it moves higher and higher into the sky, the kinetic energy is converted 
to potential energy. Eventually, when all of the kinetic energy has been converted to potential energy, the 
ball stops moving entirely and hangs in the air for a moment. Then the ball starts to fall back down, and 
the potential energy is turned into kinetic energy again. 

Just as kinetic energy and potential energy are interchangeable, work and heat are interchangeable too. 
Think of a hot-air balloon (Figure 1.21). To operate a hot-air balloon, a flame at the base of the balloon 
is used to transfer energy in the form of heat from the flame to the air molecules inside the balloon. The 
whole point of this heat transfer, though, is to capture the heat and turn it into work that causes the 
balloon to rise into the sky. The clever design of the hot-air balloon makes the conversion of heat to work 
possible. 

Lesson Summary 

• Any time we use energy, we transfer energy from one object to another. Energy can be transferred 
in one of two ways - as heat, or as work. 

• Heat is the term given to energy that is transferred from a hot object to a cooler object due to the 

53 www.ckl2.org 




Figure 1.21: A hot air balloon transfers energy in the form of heat from the flame to the air particles in 
the balloon. The design of the hot air balloon takes this energy and changes it from heat to work. 

difference in their temperatures. 

• Work is the term given to energy that is transferred as a result of a force applied over a distance. 

• Energy comes in two fundamentally different forms - kinetic energy and potential energy. 

• Kinetic energy is the energy of motion. 

• Potential energy is stored energy that depends on the position of an object relative to another object. 

• Chemical potential energy is a special type of potential energy that depends on the positions of 
different atoms and molecules relative to one another. Chemical potential energy can also be thought 
of as depending on chemical composition. 

• Energy can be converted from one form to another. 

• The total amount of mass and energy in the universe is conserved. 

Review Questions 

1 . Classify each of the following as energy primarily transferred as heat or energy primarily transferred 
as work: 

(a) The energy transferred from your body to a shopping cart as you push the shopping cart down 
the aisle. 

(b) The energy transferred from a wave to your board when you go surfing. 

(c) The energy transferred from the flames to your hotdog when you cook your hotdog over a 
campfire. 

2. Decide whether each of the following statements is true or false: 

(a) When heat is transferred to an object, the object cools down. 

(b) Any time you raise the temperature of an object, you have done work. 

(c) Any time you move an object by applying force, you have done work. 

(d) Any time you apply force to an object, you have done work. 

3. Rank the following scenarios in order of increasing work: 
www.ckl2.org 54 



(a) You apply 100 N of force to a boulder and successfully move it by 2 m. 

(b) You apply 100 N of force to a boulder and successfully move it by 1 m. 

(c) You apply 200 N of force to a boulder and successfully move it by 2 m. 

(d) You apply 200 N of force to a boulder but cannot move the boulder. 

4. In science, a vacuum is denned as space that contains absolutely no matter (no molecules, no atoms, 
etc.) Can energy be transferred as heat through a vacuum? Why or why not? 

5. Classify each of the following energies as kinetic energy or potential energy: 

(a) The energy in a chocolate bar. 

(b) The energy of rushing water used to turn a turbine or a water wheel. 

(c) The energy of a skater gliding on the ice. 

(d) The energy in a stretched rubber band. 

6. Decide which of the following objects has more kinetic energy: 

(a) A 200 lb. man running at 6 mph or a 200 lb. man running at 3 mph. 

(b) A 200 lb. man running at 7 mph or a 150 lb. man running at 7 mph. 

(c) A 400 lb. man running at 5 mph or a 150 lb. man running at 3 mph. 

7. A car and a truck are traveling along the highway at the same speed. 

(a) If the car weighs 1500 kg and the truck weighs 2500 kg, which has more kinetic energy, the car 
or the truck? 

(b) Both the car and the truck convert the potential energy stored in gasoline into the kinetic energy 
of motion. Which do you think uses more gas to travel the same distance, the car or the truck? 

8. You mix two chemicals in a beaker and notice that as the chemicals react, the beaker becomes 
noticeably colder. Which chemicals have more chemical potential energy, those present at the start 
of the reaction or those present at the end of the reaction? 

Vocabulary 

heat Energy that is transferred from one object to another object due to a difference in temperature. 
Heat naturally flows from a hot object to a cooler object. 

force Any push or pull. 

work A force applied over a distance. 

kinetic energy Energy associated with motion. 

potential energy Stored energy. Potential energy depends on an object's position or mixture's compo- 
sition. 

chemical potential energy Potential energy stored in the atoms, molecules, and bonds of matter. 

Law of Conservation of Energy Energy cannot be created or destroyed; it can only be changed from 
one form to another. 

Law of Conservation of Mass and Energy The total amount of mass and energy in the universe is 
conserved. 

55 www.ckl2.org 



Image Sources 



(1) US National Archives and Records Administration. A pony express rider, circa 1861.. Public 
Domain. 



(2 
(3 
(4 
(5 

(6 
(7 
(8 
(9 
(10 

(11 
(12 

(13 

(14 

(15 
(16 
(17 
(18 
(19 
(20 
(21 



CK-12, Richard Parsons. Shopping list chiseled on a rock.. CC-BY-SA. 

US Dept. of Health and Human Services. Receiving a vaccination.. Public Domain. 

A hot air balloon. CC-BY-SA. 

CK-12 Foundation. With no large mass nearby, this spaceship is weightless.. CC-BY-SA. 
http : //en . wikipedia . org/wiki/Image : Tehachapi_wind_f arm_2 . jpg. GNU-FDL. 

A modern cell phone.. Public Domain. 

Sharon Bewick. Some common molecules and the atoms that they are made from.. CC-BY-SA. 
http : //www . f lickr . com/photos/theyoungthousands/1978086180/. CC-BY. 

The Periodic Table.. CC-SA. 

The type of balance a chemist might use to measure the mass of an object.. Public Domain. 

Antoine Lavoisier, "The Father of Modern Chemistry.". Public Domain. 

http : //en. wikipedia. org/wiki/Image : NGC_^l^._%28NASA-med%29 . jpg Everything from an ant 
to an entire galaxy is composed of matter.. GNU-FDL, Public Domain. 

Unknown. http://commons.wikimedia.Org/wiki/File:Wagon_train.jpg. Public Domain. 

http : //en . wikipedia . org/wiki/Image : 3D_model_hydrogen_bonds_in_water .jpg. GNU-FDL. 

Maggie Black's den medeltida kokboken. Physician letting blood from a patient.. Public Domain. 

A photo of a wooden model of a Greek ship that has both sails and oars.. GNU-FDL. 
Bryan. A modern jetliner.. CC-BY-SA. 
http: //commons, wikimedia. org/wiki/File: Jiajing.jpg. Public Domain. 

The first horse-drawn street car in Seattle, Washington in 1884-- Public Domain. 



www.ckl2.org 56 



Chapter 2 

Chemistry - A Physical Science 



2.1 Measurements in Chemistry 

Lesson Objectives 



Define qualitative and quantitative observations. 

Distinguish between qualitative and quantitative observations. 

Use Quantitative observations in measurements. 

State the different measurement systems used in chemistry. 

State the different prefixes used in the metric system. 

Do unit conversions. 

Use scientific notation and significant figures. 

Use basic calculations and dimensional analysis. 

Use mathematical equations in chemistry. 



Introduction 



One of the steps in the scientific method is observation. Observation involves recording data about the phe- 
nomenon we wish to investigate. There are two different types of observations which are called qualitative 
and quantitative. Qualitative observations are those involving words only while quantitative observations 
are those involving both words and numbers. Although all the observations we can make on a phenomenon 
are valuable, quantitative observations are more helpful than qualitative. Qualitative observations are 
somewhat more vague in nature because they involve comparative terms. 

For example, a qualitative observation would be "The attendance clerk is a small woman." If the observer 
was 6 feet 4 inches tall, he/she might refer to a woman who is 5 feet 8 inches tall as "small". But if the 
observer reported this observation to a person who was 5 feet 2 inches tall, the listener would not acquire 
a good idea of the height of the attendance clerk because they would not think that a woman who is 
5 feet 8 inches tall was small (Figure ??). 



57 www.ckl2.org 



My, that 
woman is so 
tall! 




The description "a small woman" could refer to any woman whose height was between 2 feet and 6 feet 
tall depending on who did the observing. Similarly, "a small car" could refer to anything from a compact 
car to a child's toy car. The word small is a comparative term. The same is true of words like tall, fast, 
slow, hot, and cold. These words do not have exact meanings. 

Quantitative observations on the other hand, have numbers and units associated with them and are, 
therefore, more exact (Table 2.1). Even if the number is only an estimate, it is more valuable than no 
number at all. 

Table 2.1: Qualitative and Quantitative Observations Comparisons 



Qualitative (words only) 



Quantitative (words and numbers) 



The girl has very little money. 
The man was short. 
Use a small test tube. 
It is a short walk to my house. 



The girl has 85 cents. 
The man was 5 feet2 inches tall. 
Use a test tube than 12 cm long. 
It is about 1 mile to my house. 



You can see that even if the number is an estimate, a quantitative observation contains more information 
because of the number associated with the observation. (Some people might not think that a walk of one 
mile was short even though the speaker in the above case did. If an actual measuring instrument is not 
available, the observer should always try to estimate a measurement so the observation will have a number 
associated with it. 

While estimated measurements may not be accurate, they are valuable because they establish an approx- 
imate size for observations. The observation, "The car is small", provides us with certain information. 
We know that the object is some kind of automobile (perhaps real, perhaps a toy) and we know that it is 
probably smaller than a limousine because almost no one would describe a limousine as "small". Suppose 
instead, the observation had been, "The car is about 3 feet tall, 3 feet long, and 2 feet wide." While these 
estimated measurements cannot be considered to be accurate, we now know that we are not dealing with 
a compact automobile nor are we dealing with a toy car like Hot Wheels. With these estimated measure- 



www. ckl2.org 



58 



merits, we know we are dealing with a car that is about the size of a tricycle. If we discover later that 
the car was actually 2 feet high instead of 3 feet high, it is not a problem because we knew the original 
observation was an estimate since it contained the word "about". Estimates are excellent observations if 
we do not have the ability to actually measure the object. Estimated measurements qualify as quantitative 
observations. 

Here is some information you may find helpful in making estimated observations. The distance from the 
top of your index finger to the first knuckle is about one inch. The entire index finger is about three inches 
long. Your foot is probably between eight and twelve inches long. Your height is probably between five 
and six feet. The distance from your wrist to your elbow is about one foot. A twenty-five cent coin is 
about one inch across and dollar bills are about 2.5 inches wide and about six inches long (Figure 2.1). 




Figure 2.1: Helpful hints when making estimates. 

In science, accurate quantitative observations are a great deal more useful than qualitative observations. 
When we speak about accurate quantitative measurements, we are essentially referring to measurements. 
Accurate measurements are vital to science. There are many measurement systems used in the world but 
only one that is used consistently in science. That system is called the International System of Units and 
is abbreviated as SI units. You are probably already familiar with part of the SI system because part of 
the SI system is called the Metric System. 

Mass and Its SI Unit 

When you step on a bathroom scale, you are most likely thinking that about determining your weight, 
right? You probably aren't wondering if you have gained mass. Is it okay then to use either term? 

Although we often use mass and weight interchangeably, each one has a specific definition and usage. The 
mass of an object does not change; whether the object is on the earth's equator, on top of Mt. Everest, 
or in outer space, the mass will always be the same. Because mass measures how much matter the object 
contains, it has to be a constant value. 

Weight, on the other hand, is a measure of the force with which an object is attracted to the earth or body 
upon which it is situated. Since the force of gravity is not the same at every point on the earth's surface, 
the weight of an object is not constant. For example, an object weighing 1.00000 lb in Panama weighs 
1.00412 lb in Iceland. For large objects this difference may not be significant. However, since we will often 
be working with extremely tiny pieces of matter - atoms, molecules, etc. - we need to use mass and not 
weight. 

The basic unit of mass in the International System of Units (SI comes from the French name, Systeme 
Internationale) is the gram. A gram is a relatively small measurement compared to, for instance, one 
pound. 454 grams equals one pound. While pounds are helpful in measuring the mass of a package that 
needs to be mailed, grams are much more useful in science. 

One gram is equal to 1,000 milligrams or 0.001 kilogram; there are numerous intermediate measurements 
between each of these mass units as well as ones that are even larger and smaller that may be appropriate 
to the application at hand. These will be discussed in more detail in a later section. 

59 www.ckl2.org 



Length and its SI Unit 

When the four minute mile was achieved on May 6, 1954 by Roger Bannister, it was an international 
sensation. Today, many runners have broken that record. Only a few countries measure length or distance 
using miles, feet or inches. For instance, if you live in the US, you probably know your height in feet and 
inches, right? Or, if there is a mountain or even a hill near where you live, you probably know its height in 
feet. And when you discuss how far school is from your home, you probably try to figure out the distance 
in miles. 

However, most of the world measures distances in meters and kilometers; for shorter lengths, millimeters 
and centimeters will be used. For a student in Germany, she will state how many kilometers her school 
is from home, and the height of the mountain she is thinking of climbing will be given in meters (Figure 
2.2). 




Figure 2.2: Using a laser to make a precise distance measurement. 

Because the metric system is a decimal system, changing between the various measurements simply becomes 
a matter of moving a decimal point to the right or left. We will discuss this in greater depth later on. 

While a gram is a relatively small measurement, a meter is quite similar in length to a measurement with 
which we are familiar, a yard. Specifically 1 meter is equal to 1.1 yards or 39.4 inches ; one kilometer is 
equal to 0.621 miles; and one inch is equal to 2.54 centimeters. 

Volume: A Derived Unit 

Volume is used to measure how much space an object takes up. It is a derived unit, meaning it is based on 
another basic SI unit- in this case, the meter (length) was used to measure the sides of a cube, designating 
a certain volume. This volume was determined to be a cubic meter, m 3 , which is used as the standard SI 
unit of volume. This is a very large unit, and it is not very useful for most measurements in chemistry. 
A more common unit is the liter (L), which is equal to 1/1000 of a cubic meter. Another commonly used 
volume measurement is the milliliter; 1000 L = 1 L. 

One liter is the volume of the soda bottle that you might have recently purchased and have sitting in the 
refrigerator at home. You might also have a quart of milk in your refrigerator. Even though the size of the 
liter container and the milk carton may not appear to be the same, they are, in fact, almost exactly the 



www.ckl2.org 



60 



same volume. A quart is just slightly smaller in volume than a liter (1 L = 1.057 quarts). It's only the 
packaging that is different! 



Measuring Temperature 

In order to discuss temperature scales, let's briefly compare the concepts of heat and temperature. Heat 
is a measurement of the total amount of kinetic energy while temperature describes the intensity of the 
heat, or what is often referred to as the average kinetic energy of the material. When we are measuring 
the temperature of an object we are measuring its average kinetic energy. For that, we use the Celsius and 
kelvin scales. Scientists do not usually use the Fahrenheit scale. The size of a degree in kelvin is the same 
as 1 degree Celsius. The difference is that the kelvin scale begins with an absolute zero, the temperature 
at which all motion stops. To convert between the two scales you can use: K =° C + 273. Therefore, on 
the kelvin scale, water freezes at 273 K and boils at 373 K. 

You might want to make note of the following: while most mathematical calculations in chemistry require 
you to convert Celsius temperatures into kelvin, when you are given a difference in temperature, AT , you 
do not need to convert it to kelvin! A difference in temperature is the same whether it is in Celsius or 
kelvin. 



Lesson Summary 

• The International System unit for mass if the gram. 

• The International System unit for length is the meter. 

• The unit for volume is derived from a cube that is 1.00 meter on each side; therefore the volume unit 
is cubic meters. A more common unit is the liter which is j^ of a cubic meter. 

• The SI uses both °C and absolute temperature in Kelvin (K) for temperature units. K =° C + 273. 

Review Questions 

1. What are the basic units of measurement in the metric system for length and mass? 

2. What unit is used to measure volume? How is this unit related to the measurement of length? 

3. Explain the difference between weight and mass. 

4. Give both the Celsius and Kelvin values for the boiling and freezing points of water. 

5. How do you convert from Celsius to Kelvin? How does one degree Celsius compare with one Kelvin? 

6. If someone told you that a swimming pool's temperature was 275 K, would it be safe for you to go 
for a swim? 

7. Determine which metric measurement you would use for each of the following: 

(a) The distance to the moon. 

(b) The mass of a donut. 

(c) The volume of a drinking glass. 

(d) The length of your little finger. 

Further Reading / Supplemental Links 

• Website with lessons, worksheets, and quizzes on various high school chemistry topics. Lesson 1-3 is 
on Measuring Matter, http://www.fordhamprep.org/gcurran/sho/sho/lessons/lessonl3.htm 

61 www.ckl2.org 



Vocabulary 

International System of Units, SI The SI system of units is the modern form of the metric system 
and is generally a system devised around the convenience of multiples of 10. 

Kelvin temperature scale The kelvin is unit of increment of temperature and is one of the seven SI 
basic units. The Kelvin scale is thermodynamic absolute temperature scale where absolute zero is 
the absence of all thermal zero. At K = 0, there is no molecular motion. The kelvin is not referred 
to as a "degree", it is written simply as K, not ° K. 



2.2 Using Measurements 

Lesson Objectives 

• Understand the metric system and its units. 

• Convert between units. 

• Use scientific notation in writing measurements and in calculations. 

• Use significant figures in measurements. 



Introduction 

The metric system is a decimal system. This means that making conversions between different units of 
the metric system are always done with factors of ten. Let's consider the English system - that is, the 
one that is in everyday use in the US as well as England - to explain why the metric system is so much 
easier to manipulate. For instance, if you need to know how many inches are in a foot, you only need to 
remember what you at one time memorized: 12 inches = 1 foot. But now you need to know how many 
feet are in a mile. What happens if you never memorized this fact? Of course you can look it up online or 
elsewhere, but the point is that this fact must be given to you as there is no way for you to derive it out 
yourself. This is true about all parts of the English system: you have to memorize all the facts that are 
needed for different measurements. 



The Metric System 

In the metric system, you need to know (or yes, memorize) one set of prefixes and then apply them to each 
type of measurement. Then if a larger measurement is needed, such as kilometers, but you have used a 
meter stick, you only need to move the decimal to convert the units. 

Example: If you have measured the distance as 60.7 meters, what is the length in kilometers? 

Solution: 60.7 meters = 0.0607 kilometers since there are 1,000 meters in 1 kilometer. 

Not only can you easily convert from kilometers to meters, but conversions, such as liters to cubic meters, 
are also easy. Try converting from cubic feet to gallons! All metric system conversions simply require the 
moving of the decimal and/or adding zeros. You don't even need a calculator. On the other hand, if you 
had to convert from miles to inches, not only would you have to remember all of the conversion factors, 
but you would probably also need a calculator to make the conversion. 

www.ckl2.org 62 



Metric Prefixes 

The metric system uses a number of prefixes along with the base unit. To review: the basic unit of mass 
is a gram (g), that of length is meter (m), and that of volume is liter (L). When the prefix centi is place in 
front of gram, as in centigram, the measure is now y^ of a gram. When milli is placed in front of meter, 
as in millimeter, the measure is now y^ of a meter. Common prefixes are in Table 2.2: 

Table 2.2: 

Prefix Meaning Symbol 

pico- 1CT 12 p 

nano- 10~ 9 n 

micro- 10 -6 fi 

milli- 10~ 3 m 

centi- 10~ 2 c 

deci- 1CT 1 d 

kilo- 10 3 k 



Unit Conversions 

Making conversions in the metric system is relatively easy: you just need to remember that everything is 
based on factors of ten. For example, let's say you want to convert 0.0856 meters into millimeters. Looking 
at the chart above, you can see that 1 millimeter is 10 -3 meters; another way to say this is that there are 
1000 millimeters in one meter. You can set up a mathematical expression as follows 

(0.0856 m) X (1000 mm/1 m) =? mm 

When you solve this equation, you first want to see which units to divide out. In this case, you notice 
meters appear in both the numerator and denominator, so you will be able to cancel them. 

(0.0856 m) X (1000 mm/1 m) =? mm 

Now all that is left to do is multiply 0.0856 by 1000. To do this, you are just going to move the decimal 
point three places to the right: 

(0.0856 m) X (1000 mm/1 m) = 85.6 mm 

Example: Convert 153 grams to centigrams. 

Solution: 153 grams X cen lgram& — 15300 centigrams 

Scientific Notation 

Scientific notation is a way to write very large or very small numbers (Figure 2.3); the decimal point is 
moved so that there is one digit in the unit's position and all of the decimal places are held as a power 
of ten. This is important in chemistry because many of the measurements we make either involve very 
large numbers of atoms/molecules or very tiny measurements, such as masses of electrons or protons. For 
example, consider a number such as 839,000,000. While this number is not too difficult to write out, it 
is more conveniently written in scientific notation. Written in scientific notation, this number becomes: 
8.39 x 10 8 . The "10 8 " means that ten is multiplied by itself eight times: 10 x 10 x 10 x 10 x 10 x 10 x 10 x 10. 
As you can see, writing 10 8 is much easier! 

We can also use scientific notation to write very small numbers. Take a number such as 0.00000481. It is 



63 www.ckl2.org 




Figure 2.3: Even using large distance units such as kilometers, you would still need to use scientific notation 
to measure the size of this galaxy. 

easy to make mistakes in counting the number of zeros in this number. Also, many calculators only let you 
enter in a certain number of digits. When we write this in scientific notation, it is important to notice that 
the measurement is less than one, therefore, the exponent on 10 will be negative: this number becomes 
4.81 x 10~ 6 . In this case, the decimal point was moved six places to the right. 

It is important that you know how to perform calculations using numbers written in scientific notation. 
For example, the following problem shows two numbers with exponents being multiplied together: 

(2.90xl0 3 )(1.60xl0 6 ) =? 

To solve this problem, you would multiply the terms (2.90 and 1.60) like you normally would; then you 
would add the exponents: 

2.90x1.60 = 4.64 
10 3 x 10 6 = 10 9 

Therefore, combining these values gives the answer 4.64 x 10 9 . 

Significant Figures 

The tool that you use determines the number of digits that will be in a measurement. For example, if you 
say an object has a mass of "5 kg", that is not the same as saying it has a mass of "5.00 kg" since you 
must have measured the masses with two different tools - the two zeros in "5.00 kg" would not be written 
if the tool that was used could not measure to two decimal places. Even though the mass seems to be the 
same, the uncertainty of the measurement is not. When you say "5 kg," that means you have measured 
the mass to within +/ - 1 kg. The actual mass could be 4 or 6 kg. For the 5.00 kg measurement, you have 
measured the mass to within +/ - 0.01 kg, so the actual mass is between 4.99 and 5.01 kg. 

Using Significant Figures in Measurements 

How do you know how many significant figures are in a measurement? General guidelines are as follows: 
www.ckl2.org 64 



• Any nonzero digit is significant [4.33 has three significant figures]. 

• A zero that is between two nonzero digits is significant [4.03 has three significant figures]. 

• All zeros to the left of the first nonzero digit are not significant [0.00433 has three significant figures] . 

• Zeros that occur after the decimal are significant. [40.0 has three significant figures. The zero after 
the decimal point tells us that the value was measured to the tenths place] . 

• Zeros that occur without a decimal are not significant [4000 has one significant figure since the zeros 
are holding the 4 in the thousands position]. 

Examples: 

1) How many significant figures are in the number 1.680? 

Solution: There are three nonzero digits and one zero appears after the decimal point. Therefore, there 
are four significant figures. 

2) How many significant figures are in the number 0.0058201? 

Solution: There are 4 nonzero digits and 1 zero between two numbers. Therefore, there are 5 significant 
figures. The first three zeros are not significant since they are simply holding the number away from the 
decimal point. 

Lesson Summary 

• The metric system is a decimal system; all magnitude differences in units are multiples of 10. 

• Unit conversions involve creating a conversion factor. 

• Very large and very small numbers are expressed in exponential notation. 

• Significant figures are used to express uncertainty in measurements. 

Review Questions 

1. Convert the following linear measurements: 

(a) 0.01866 m = cm 

(b) 2156 mm = m 

(c) 15.38 km = m 

(d) 1250.2 m = km 

2. Convert the following mass measurements: 

(a) 155.13 mg — kg 

(b) 0.233 g = mg 

(c) 1.669 kg = g 

(d) 0.2885 g = mg 

3. Write the following numbers in scientific notation: 

(a) 0.0000479 

(b) 251,000,000 

(c) 4260 

(d) 0.00206 

4. How many significant figures are in the following numbers? 

(a) 0.006258 

(b) 1.00 

(c) 1.01005 

(d) 12500 

65 www.ckl2.org 



Further Reading / Supplemental Links 

The website listed below offers lessons, worksheets, and quizzes on many topics in high school chemistry. 

• Lesson 1-7 is on Temperature Conversion. 

• Lesson 2-1 is on the International System of Measurements. 

• Lesson 2-3 is on Significant Figures. 

• Lesson 2-5 is on Scientific Notation, http://www.fordhamprep.org/gcurran/sho/sho/lessons/ 
lessonl2.htm 

Vocabulary 

scientific notation A shorthand way of writing very large or very small numbers. The notation consists 
of a decimal number between 1 and 10 multiplied by an integral power of 10. It is also known as 
exponential notation. 

significant figures Any digit of a number that is known with certainty plus one uncertain digit. Begin- 
ning zeros and placeholder zeros are not significant figures. 

2.3 Using Mathematics in Chemistry 

Lesson Objectives 

• Use units in problem solving. 

• Do problem solving using dimensional analysis. 

• Use significant figures in calculations. 

Introduction 

Unit terms are the words following a measurement that tell you on which standard the measurement 
is based. Every measurement must have a unit term. The unit terms also follow the algebraic rules 
of exponents and cancellation. Carrying the unit terms through mathematical operations provide an 
indication as to whether the mathematical operation was carried out correctly. If the unit term of the 
answer is not correct, it is an indication that the mathematical operation was not done correctly. 

Using Units in Problem Solving 

Anytime we have to do a calculation, it is important to include the units along with the actual numbers. 
One reason is that you can often decide how to solve the problem simply by looking at the units. For 
example, let's say you are trying to calculate solubility. One of the units used for solubility is grams/liter 
(g/L). Imagine that you have forgotten how to do the mathematical calculation for this problem, but you 
have measured how many grams of a solid dissolved into a certain number of liters. Looking at the units 
of the values that you have (g and L) and at the units of the answer you want to get (g/L), you can figure 
out the mathematical set-up. The g/L unit allows you to know it needs to be "grams divided by liters." 

You will also note that as you do a calculation, you will be working with units in a similar manner as you 
would a number. Just as with numbers, units can be divided out when that specific unit appears in the 
numerator as well as the denominator. 

www.ckl2.org 66 



As a final note on units, think of them in an "apples and oranges" context. You can't subtract meters from 
kilometers without first converting the measurements into common units. Always check a measurement's 
units to make sure that they are appropriate for a given calculation. 

Using Conversion Factors 

Conversion factors are used to convert one unit of measurement into another. A simple conversion factor 
can be used to convert meters into centimeters, or a more complex one can be used to convert miles per 
hour into meters per second. Since most calculations require measurements to be in certain units, you will 
find many uses for conversion factors. What always must be remembered is that a conversion factor has to 
represent a fact; this fact can either be simple or much more complex. For instance, you already know that 
12 eggs equal 1 dozen. A more complex fact is that the speed of light is 1.86 x 10 5 miles/sec. Either one 
of these can be used as a conversion factor depending on what type of calculation you might be working 
with. The following section provides you with more examples of this. 

Dimensional Analysis 

When using conversion factors (and for that matter, a lot of other calculations), a process called dimensional 
analysis is extremely useful. Dimensional analysis allows you to make a number of unit conversions in a 
single calculation. It will also help you keep the units straight. 

Example: A car travels 58.5 miles, using 1.5 gallons of gasoline. How do you express this in kilometers/liter? 
You know that there are 3.78 liters in a gallon, and a kilometer is 0.62 miles. How would you make this 
conversion? 

Solution: 

You first need to write out a mathematical expression showing all your conversion factors and units: 



58.5 miles 
1.5 gallons 



)/ 1 gallon \ /l kilometer \ 
X \3.78 liters/ X \ 0.62 miles J 



Next, you need to check for units to divide out: 

58.5 miles 



1.5 gallons 



)/ 1 gallon \ /l kilometer \ 
X \3.78 liters/ X \ 0.62 miles / 



Notice that at this point you are left with kilometers in the numerator and liters in the denominator. Your 
last step is to multiply your numbers, and your answer will be in kilometers/liter: 

'58. 5\ / 1 \ /l kilometer\ /16. 64106503 kilometers\ 
x 



(ttH 



3.78 liters/ \ 0.62 / \ liter / 

This is the answer that your calculator will give to you. However it is not the correct answer. For that we 
have to proceed to the next section: 

Using Significant Figures in Multiplication and Division 

Whenever we do a calculation, we need to pay attention to the significant figures. The rule is that your 
final answer can only be as precise as your least precise measurement. This means that the least precise 
tool used for any measurement in the calculation will determine how precise the answer will be. 

67 www.ckl2.org 



For multiplication and division, first determine the number of significant figures in each of the measure- 
ments; the number of significant figures in your answer will be the same as the least number in the 
calculation. For example, if you multiplied the number 1.02584 by 2.1, your answer can only have two 
significant figures. The same rule applies for division. 

Example: Divide the number 125.688 by 14.01. Express your answer using correct significant figures. 

Solution: 125.688 has 6 significant figures, and 14.01 has 4. Therefore, your answer can only have 4 
significant figures. 

14.01 

An important point to remember is that if you are multiplying or dividing by an exact number, then you 
treat that number as having an infinite number of significant figures. An exact number is a number that 
is written without all of its known significant figures. For instance, one meter or one dozen have many 
significant figures but we just don't write all of them, so these kind of measurements never determine the 
number of significant figures in a calculation. There is another type of exact number and that is using 
a measurement such as five people, ten dogs, or one cat. Again, these do not determine the number of 
significant figures in an answer. 

Example: You have one dozen identical objects, with a total mass of 46.011 grams. What is their average 
mass? 

Solution: 

12 = 3.83425, which rounds off to 3.8342 (5 significant figures, the same as 46.011) 

Using Significant Figures in Addition and Subtraction 

There is a different rule for determining significant figures when adding or subtracting measurements. Now, 
you will need to look for the measurement with the least number of significant figures to the right of the 
decimal place; this number of decimal places will determine the number of significant figures to be used in 
the answer. 

Example: What is the sum of 14.3 and 12.887? 

Solution: 

14.3 + 12.887 = 27.187 
The number 14.3 only has 1 digit to the right of the decimal point, so our answer is rounded off to 27.2. 

Lesson Summary 

• Dimensional analysis aids in problem solving. 

• Conversion factors are created by unit analysis. 

• Significant figures must be carried through mathematical operations. 

• The answer for an addition or subtraction problem must have digits no further to the right than the 
shortest addend. 

• The answer for a multiplication or division problem must have the same number of significant figures 
as the factor with the fewest significant figures. 

www.ckl2.org 68 



Review Questions 

1. Perform the following calculations and give your answer with the correct number of significant figures: 

(a) 0.1886x12 

(\y) 2.995 
\ u ) 0.16685 

(c) -&$- 

VW 0.1223 

(d) 910x0.18945 

2. Perform the following calculations and give your answer with the correct number of significant figures: 

(a) 10.5 + 11.62 

(b) 0.01223 + 1.01 

(c) 19.85-0.0113 

3. Do the following calculations without a calculator: 

(a) (2.0xl0 3 )(3.0xl0 4 ) 

(b) (5.0xl0- 5 )(5.0xl0 8 ) 

(c) (e.oxio^xr.oxio- 4 ) 

,,n (3.0xlQ- 4 )(2.0xlQ- 4 ) 
W 2.0X10- 6 

Further Reading / Supplemental Links 

The website listed below offers lessons, worksheets, and quizzes on many topics in high school chemistry. 

• Lesson 2-4 is on the Factor-Label Method of Unit Conversion, http://www.fordhamprep.org/ 
gcurran/sho/sho/lessons/lessonl2 . htm 

Vocabulary 

dimensional analysis A technique that involves the study of the dimensions (units) of physical quanti- 
ties. It affords a convenient means of checking mathematical equations. 

2.4 Using Algebra in Chemistry 

Lesson Objectives 

• Be able to rearrange mathematical formulas for a specific variable. 

• Have an understanding of how to use units in formulas. 

• Be able to express answers in significant figures and with units. 

Introduction 

During your studies of chemistry (and physics also), you will note that mathematical equations are used 
in a number of different applications. Many of these equations have a number of different variables with 
which you will need to work. You should also note that these equations will often require you to use 
measurements with their units. Algebra skills become very important here! 

69 www.ckl2.org 



Solving Formulas with Algebra 

Sometimes, you will have to rearrange an equation to get it in the form that you need. When you are 
working with an equation such as D = y (density = mass/volume) and asked to solve for density, it is 
relatively easy; all you have to do is substitute the measurements and solve - of course, keeping in mind 
significant figures! 

If you are asked to solve the above equation for M, then you will need to manipulate the equation to isolate 
the desired variable, in this case in the form of "M =." To do this, you will need to move the V from the 
right side to the left side of the equation. As the V is in the denominator, you will need to multiply both 
sides of the equation by V: 

VXD=(V)(£) 

Multiplying this out: 

VxD = M 

A similar process is used if you need to solve for V: 

Take the previous equation (V x D = M) and divide both sides by D: 

Solving, this becomes: 

M 

What if you are given a more complex equation, like PV = nRT, and asked to solve for "«"? You need to 
follow the same steps as you did in the above examples. The only difference is that there are more symbols 
to rearrange. 

Solution: 

Look at the original equation: PV — nRT . Our goal is to get "n" on one side of the equation by itself. 

To remove the RT from the right side, we will divide both sides by RT: 

PV nRT 



RT RT 

After solving, we are left with: 

PV 

n — 

RT 

Algebra with Units and Significant Figures 

So far, you've learned about units, significant figures, and algebraic manipulation of equations. Now it's 
time to put all three of these together. We'll start with a simple example: density. Density is a measure 
of the amount of mass per unit of volume, and a common unit used is g/niL. In the first example, we're 
going to do a straightforward calculation of density from a given mass and volume. 

www.ckl2.org TO 



Example 1 

What is the density of an object that has a mass of 13.5 g and a volume of 7.2 mL? 

Solution: 

The equation is as follows: 

mass 
D 



volume 



Substituting in the known values (with units): D = 72m L 

Finally, solving the equation and rounding off the answer based in significant figures: 

13 5 e 

D = = 1.9g/mL(2 significant figures) 

7.2 mL b/ v s s i 

This calculation was easy to do because there was no rearranging of the equation and no cancellation of 
units. 

For the next example, let's look at a more complex calculation. 

Example 2: 

A sample of an ideal gas has a volume of 14.2 L at a pressure of 1.2 atm. If the gas pressure is increased 
to 1.8 atm, what is the new volume? 

Solution: 

This problem uses Boyle's law: 

Pi Vi = P 2 V 2 

All the variables are known except for V 2 , so the equation needs to be rearranged to solve for the one 
unknown. We can do this by dividing both sides by P 2 : 

P1V1 

Now we substitute in the known values and their units: 

(1.2 atm) (14.2 L) 

(1.8 atm) 

Next, we cancel out units: 

(1.2)(14.2 L) 

Finally, we calculate our answer and round off to the appropriate number of significant figures: 

V 2 = 9.5 L 

Lesson Summary 

• Students of chemistry need to be able to use algebra in their calculations. 

71 www.ckl2.org 



V 2 



V 2 



Review Questions 

1. For the equation PV = nRT, re- write it so that it is in the form of " T =." 

2. The equation for density is D = M/V. If D is 12.8 g/cm , and M is 46.1 g, solve for V, keeping 
significant figures in mind. 

3. The equation P\ V\ = Pi V2, known as Boyle's law, shows that gas pressure is inversely proportional 
to its volume. Re-write Boyle's law so it is in the form of V\ =?. 

4. The density of a certain solid is measured and found to be 12.68 g/mL. Convert this measurement 
into kg/L. 

5. In a nuclear chemistry experiment, an alpha particle is found to have a velocity of 14,285 m/s. 
Convert this measurement into miles/hour. 

Further Reading / Supplemental Links 

• College Chemistry- Schaum's Outlines. Jerome Rosenberg and Lawrence Epstein, McGraw-Hill, 
1997. 

• Chemistry, 7 th Edition, Raymond Chang, McGraw-Hill, 2002. 

• Chemistry, J. Dudley Herron, Jerry L. Sarquis, Clifford L. Schrader, David V. Frank, Mickey Sarquis, 
David A. Kukla. D.C. Heath and Co., 1993. 

• World of Chemistry, Steven S. Zumdahl, Susan L. Zumdahl, and Donald J. DeCoste. McDougal 
Littell, 2007. 

Labs and Demonstrations for Using Algebra in Chemistry 
Density Determination 

Pre-Lab Discussion 

Density is defined as the mass per unit volume of a substance. The table below lists the densities of some 
well known substances. 

Table 2.3: Density of Some Common Substances 
Substance Density Substance Density 



Water 


1.0 g/cm 3 


Aluminum 


2.7 g/cm 3 


Oxygen gas 


0.0013 g/cm 3 


Iron 


8.9 g/cm 3 


Sugar 


1.6 g/cm 3 


Lead 


11.3 g/cm 3 


Table salt 


2.2 g/cm 3 


Gold 


19.3 g/cm 3 



Density measurements allow scientists to compare the masses of equal volumes of substances. If you had 
a piece of lead as large as your fist and a piece of gold as large as your thumb, you would not know which 
substance was innately heavier because the size of the pieces are different. Determining the density of the 
substances would allow you to compare the masses of the same volume of each substance. The process for 
finding the density of a substance involves measuring the volume and the mass of a sample of the substance 
and then calculating density using the following formula. 

mass in grams 

Density — — 

volume in mL 

Example: Calculate the density of a piece of lead whose mass is 226 grams and whose volume is 20.0 mL. 
www.ckl2.org 72 



Also calculate the density of a sample of gold whose mass is 57.9 grams and whose volume is 3.00 mL. 
Solution: 



Density of Lead 
Density of Gold 



mass 226 g 

volume 20.0 mL 

mass 57.9 g 

volume 3.00 mL 



11.3 g/mL 
19.3 g/mL 



Methods of Measuring Mass and Volume 

The mass of substances is measured with a balance. In the case of a solid object that will not react with 
the balance pan, the object may be placed directly on the balance pan. In the case of liquids or reactive 
solids, the substance must be placed in a container and the container placed on the balance pan. In order 
to determine the mass of the substance, the mass of the container is determined before hand (empty) and 
then the container's mass is subtracted from the total mass to determine the mass of the substance in the 
container. There are several common procedures for determining the volume of a substance. The volume 
of a liquid is determined by pouring the liquid into a graduated cylinder and reading the bottom of the 
meniscus. For a regularly shaped solid, the volume can be calculated from various linear measurements. 



CUBE 



SPHERE 



CYLINDER 






V=(x)(y)(z) 



V=lnr 3 



V=nr 2 h 



For an irregularly shaped object, a graduated cylinder is partially filled with water and the volume mea- 
sured. The object is then submerged in the water and the new volume measured. The difference between 
the two volumes is the volume of the submerged object. 

Equipment: Specific gravity blocks, graduated cylinders (10 mL and 100 mL), thread, millimeter ruler, 
balance, distilled water, glycerol. (If you have an overflow can, it also works well for submersion.) 

Procedure: 

1. Obtain a regularly shaped object from your teacher. Measure and record its mass in grams and its 
dimensions in centimeters. 

2. Add approximately 50 mL of tap water to a 100 - mL graduate and record its exact volume. Tie a 
thread to the block and carefully immerse it in the cylinder of water. Record the new volume in the 
cylinder. 

3. Measure and record the mass of a clean, dry, 10 - mL graduated cylinder. 

4. Add exactly 10.0 mL of distilled water to the cylinder. Measure and record the combined mass of the 
cylinder and water. 

5. Repeat steps 3 and 4 with glycerol instead of water. 

Data Table 



73 



www.ckl2.org 



Object 

code name or letters 

width 

height 

length 

volume of water before block 
volume of water after block _ 
Distilled Water 

volume 

mass of empty graduate 

combined mass 

Glycerol 
volume 



mass of empty graduate 

combined mass 

Calculations 

1. Find the volume of the solid object using the dimensions and appropriate formula. 

2. Find the volume of the solid object using water displacement. 

3. Find the density of the block using the volume from calculation 1. 

4. Find the density of the block using the volume from calculation 2. 

5. Find the density of the distilled water. 

6. Find the density of the glycerol. 

7. If your teacher gives you the accepted values for the densities in this lab, calculate the percent error 
for your values. 

(experimental value) — (actual value) 

% error = — X 100 

[actual value) 



Image Sources 



(1) Using a laser to make a precise distance measurement.. CC-BY-SA. 

(2) http://www.flickr.com/photos/pingnews/474783096/. Public Domain. 

(3) Richard Parsons. Helpful hints when making estimates.. CC-BY-SA. 



www.ckl2.org 74 



Chapter 3 

Chemistry in the Laboratory 



3.1 Making Observations 

Lesson Objectives 

• Define qualitative and quantitative observations. 

• Distinguish between qualitative and quantitative observations. 

• Use Quantitative observations in measurements. 

Introduction 

Take out a piece of paper and record a chart similar to the one below. Look up from this text and scan 
the room. Write down what you see around you in as much detail as you feel necessary to remember it 
after you walk away. Better still, you want to be able to show your chart to someone who can then be able 
to picture where in the room you were sitting. A chart is usually necessary to record these observations 
simply for organization (Table 3.1). 

Table 3.1: 

Item Observation 

1. 

2. 
3. 

4. 
5. 



Science depends on keeping records of observations for later interpretations. These interpretations may 
lead to the development of theories or laws. Without accurate observations, scientists cannot make any 
interpretations and therefore cannot draw conclusions. 

Here is a simple test for you. Pretend we're visiting a forensic scientist, hired to investigate the scene of a 
crime. You are only asked to analyze the observations gathered by the other scientists at the scene. You 
must try not to make too many assumptions yet try and make your decision based on the data at hand. 
One summer evening, Scott and Brenda came home from work to find their house in shambles. Neighbors, 



75 www.ckl2.org 



friends, and colleagues are baffled by the strange occurrence. The television was found on in the house. 
Food was on the table ready to be eaten. All of Scott's coin collections, his precious metals, and Brenda's 
prized possession - her statue of Galileo - were gone. Foul play is suspected. The lead investigator gives 
you the following observations gathered from the scene and suspects (Table 3.2): 

Table 3.2: 



Observations at Scene Suspect 1: 180 lb male Suspect 2: 220 lb male Suspect 3: 120 lb female 



Blood type = B 
Fiber sample = 
polyester 

Powder found = white 
Shoe Print found = 
work boot 



Blood type = B 
Sweater = polyester 



Blood type = B 
Blazer = wool knit 



Works in sugar factory Pastry chef 



Would not comply 
Pants = polyester 

Car sales woman 



From the table, can you deduce who might have been involved in the alleged crime? Do you need more 
information? How good are the observations in order for you, the scientist, to make accurate conclusions? 
What will you base your decision on? What other information do you need? Remember someone will be 
charged for this crime. Observations are the key to science all around us and in our everyday lives. 

Qualitative Observations 

Science is full of observations but of two different types. What we see, smell, feel, and hear are observations 
that scientists depend on to determine whether chemical reactions have been occurring or have come to 
completion. This is one type of observation known as a qualitative observation. Qualitative observations 
give the descriptive properties of a substance or being and therefore are without numbers. 

When you made your table above, what kind of observations did you make? Take a look at the table. Did 
you note any colors from the surroundings? Was there a window nearby? If so, was it open? Did you 
happen to hear any sounds from outside the window? Did you see a vehicle drive by? If so, what color 
was it? Take a look at the sample question below and see if you can determine the qualitative observations 
from each of the figures in the question. 

Sample Problem: List the qualitative observations for each of the figures below. 

Left image: Fog caused by dry ice. (Source: http://commons.wikimedia.Org/wiki/File:DryIceSublimation. 
JPG. Public Domain) 

Center image: Tulips. (Source: http://c0mm0ns.wikimedia.0rg/wiki/File:Vi0lett_tulips.jpg. Pub- 
lic Domain) 

Right image: Soda. (Source:http : //en. wikipedia. org/wiki/Image: Soda_bubbles_macro.jpg. Public 
Domain) 




Solution: 



www.ckl2.org 



76 



Fog caused by dry ice: fog or smoke coming from top of the cup pouring over onto the tabletop 

Tulips: purple tulips, sky with clouds in the background 

Soda: bubbles, almost looks like effervescence from soda pop when you pour a glass of cola 

Quantitative Observations 

Sometimes qualitative measurements are enough to give an accurate representation of the events occurring. 
In other cases, scientists need more information than what the senses offer in order to make correct 
interpretations and then conclusions. Say, for example, there was a window in your classroom and outside 
the window you see a battleship gray car zoom by, speeding you suspect because behind him you hear the 
sirens of a police car. You make the interpretation that the car is speeding because of the police sirens. 
The police could, in fact, be on another call. The only real accurate qualitative observations you can make 
here are that you see a battleship gray car drive by, you hear the sirens of the police car, and you see the 
police car drive by with sirens on. Now if you say, happen to have a hand held radar gun, and could then 
measure that the car was travelling 50 mph in a 35 mph zone, then you could conclude the police were 
chasing the car. 

Quantities are a useful strand of observations. When you have observations that involve the use of numbers, 
we refer to these as quantitative observations because they have amounts. In our car chase example above, 
the measurement of 50 mph and 35 mph are both quantitative measurements. If we said that it is 85° F 
outside in British Columbia, the temperature of 85° F is a quantitative observation. Now, if we said that 
it is 85° F outside in sunny British Columbia, the temperature of 85° F is a quantitative observation, and 
the "sunny" is a qualitative observation. See how it works? Now you try. 

Sample Question: 

Pick out the quantitative and qualitative observations from each phrase: 

1. 3 g of NaCl dissolves in 10 mL of H^O to make a clear solution. 

2. The spider on the wall only has seven legs remaining but is still big and hairy. 

3. When 0.5 mL of a solution is put into a flame, the flame turns a brilliant green. 



Solution: 

1. Quantitative 

2. Quantitative 

3. Quantitative 



3 g 10 mL; qualitative: clear, 
seven; qualitative: big, hairy. 
0.5 mL; qualitative: brilliant green. 



Lesson Summary 

• Qualitative observations describe the qualities of a substance or event. 

• Qualitative observations are made using the senses (except taste - used only when appropriate). 

• Qualitative observations do not involve numbers. 

• Quantitative observations describe the quantities of a substance or event. 

• Quantitative observations use numbers in the descriptions for the substance or event. 

• Observations, either qualitative or quantitative, or both, are used as tools by scientists to make 
representations and then interpretations about the surroundings. 

Review Questions 

1. Indicate in the following chart (Table 3.3) whether the observation is qualitative or quantitative. 

77 www.ckl2.org 



Table 3.3: 



Number 



Observation 



Qualitative or Quantitative 



1. 

2. 

3. 

4. 
5. 
6. 

7. 
8. 

9. 

10. 
11. 

12. 



The temperature of this room is 

25°C. 

It is comfortably warm in this 

room. 

Most people have removed their 

coats. 

The building is 25 stories high. 

It is a very tall building. 

The building is taller than any 

nearby trees. 

The bottle is green. 

The bottle contains 250 mL of 

liquid. 

Robert bought his son a small 

car. 

The car is smaller than his hand. 

The car is about three inches 

long. 

The race is about 27 miles long. 



Vocabulary 

qualitative observations Describe the qualities of something and are described without numbers. 
quantitative observations Observations that involve the use of numbers (quantities). 

Further Reading / Supplemental Links 

• http://en.wikipedia.org/wiki 

3.2 Making Measurements 

Lesson Objectives 

• Match equipment type based on the units of measurements desired. 

• Determine significant figures of the equipment pieces chosen. 

• Define accuracy and precision. 

• Distinguish between accuracy and precision. 

Introduction 

As we learned in the previous section, qualitative observations require the use of the senses to gather data 
in order to interpret what is happening in our surroundings and then make conclusions based on these 



www.ckl2.org 



78 



interpretations. Quantitative observations gather data by using measurements. From these measurements 
we can interpret the data and draw conclusions. How exactly do scientists gather all of this numerical 
data? What kind of equipment is necessary and for what purposes? How accurate is it? Let's take a 
look, first at some of the typical equipment used in chemistry and then at the skills necessary to determine 
accuracy and precision. Let's explore the quantitative side of chemistry. 

Equipment Determines the Unit of the Measurement 

Think of the last laboratory experiment that you did. What kind of equipment did you use? If you were 
measuring out a volume of a liquid, did you use a beaker or a graduated cylinder? 

Beakers used in experiments. (Source: http://c0mm0ns.wikimedia.0rg/wiki/File:Beakers.jpg. CC- 
BY-SA) 




100 mL graduated cylinder. (Source: http://en.wikipedia.Org/wiki/Image:Graduated_cylinder. 
jpg. Public Domain) 




Look at the two figures above; if you were required to measure out 65 mL, what instrument would you most 
likely want to use? The graduated cylinder has graduations every 10 mL and then further graduations 
every 5 mL. The beakers could have graduations every 10 mL, 50 mL, or 100 mL depending on which 
type you use. It would be easier to measure out the volume in a graduated cylinder. What if, in this same 
lab, you needed to mass out 3.25 g of sodium chloride, NaCl. Look at the two figures below and determine 
which piece of equipment you would use. 

79 www.ckl2.org 



A pan balance. (Source: http : //en . wikipedia . org/wiki/File : Balance_ / C3°/„AO_tabac_1850 . JPG . GNU- 
FDL) 




A digital balance. (Source: http : //commons . wikimedia . org/wiki/File : Digi-keukenweegschaall284 . 
JPG. GNU-FDL) 



I 


'*dkt 






1 


1 


I 


p'tuuf firiVFfl 




'UniillPMLVUl 


J 








■^ 






m 




L 


^r~ 


^^^^^^^^^^— ^ 


k 




' aarw; * 




^ 




^~ 




^^"^ 











The pan balance measures only to +/ - 0.1 g. Therefore, you would have to mass out 3.3 g of NaCl rather 
than 3.25 g. The digital balance can measure to +/ - 0.01 g. With this instrument you could measure 
exactly what you need, depending on your skill of course! 

The equipment you choose also determines the units in your measurement and vice versa. For example, 
if you are given graduated cylinders, beakers, pipettes, burettes, flasks, or bottles, you are being asked to 
measure volume. Volume measurements in the International System of Units use the metric system rather 
than the imperial system in order to standardize these measurements around the globe. Thus, for volume 
measurements, we use liters (L) for large volumes and milliliters (mL) for smaller volumes measured in the 
lab. Look at the figure below and determine what volumes are present in each piece of equipment. 

Volume equipment pieces. (Source: Richard Parsons. CC-BY-SA) 



www.ckl2.org 



80 




I 



Graduated Cylinder Graduated Pipet 



The contrary is also true. What if you were to measure out 5 g of a solid, or 3 cm of wire, or the temperature 
of a solution; would any of the objects in Figure 5 be helpful? Why not? These objects are not helpful 
because these units of measurement are not volumes and all of these pieces of equipment measure volume. 
For the measurements you need to take, you would need different pieces of equipment. Look at the figure 
below and match the three required measurements with the pieces of equipment shown. 

a) 5 g of a solid 

b) 3 cm of wire 

c) temperature of a solution 

Examples of measuring devices. (Source: http : //en . wikipedia . org/wiki/ Image : Steel_ruler_closeup . 
jpg http : //en . wikipedia . org/wiki/File : Clinical_thermometer_38 . 7 . JPG http : //en . wikipedia . 
org/wiki /Image: Digi-keukenweegschaall284.jpg. GNU-FDL, GNU-FDL, GNU-FDL) 




Equipment Determines the Significant Figures 

In the previous section, we looked at a lot of equipment that is used for measuring specific units. The 
graduated cylinder that measures volume, the balance that measures mass, and the thermometer that 
measures temperature are a few that we looked at before. We also saw that of two types of balances, 
one type of balance can more precisely measure mass than the other. The difference between these two 
balances has to do with the number of significant digits that the balances are able to measure. Remember 
the pan balance could measure to +/ - 0.1 g and the digital balance can measure to +/ - 0.01 g. 

Before going any further, what do we recall about significant digits? A measurement can only be as 
accurate as the instrument that produced it. A scientist must be able to express the accuracy of a number, 
not just its numerical value. 

The instruments that we choose for the laboratory experiments depend on the required amount of accuracy. 



81 



www.ckl2.org 



For example, if you were to make a cup of hot chocolate at home using powdered cocoa, you would probably 
use a measuring spoon or a teaspoon. Compare this to the requirement of massing out 4.025 g of sodium 
bicarbonate for a reaction sequence you are doing in the lab. Would the teaspoon do? Probably not! You 
would need to have what is known as an analytical balance that measures to +/ - 0.001 g. 

Accuracy and Precision 

Accuracy and precision are two words that we hear a lot in science, in math, and in other everyday events. 
They are also, surprisingly, two words that are often misused. How often have you heard these terms? 
For example, you often hear car advertisements that talk about their precision driving ability. But what 
do these two words mean. Accuracy is how close a number is to the actual or predicted value. If the 
weatherperson predicts that the temperature on July 1 st will be 30°C and it is actually 29°C, she is likely 
to be considered pretty accurate for that day. 

Once you have gone into the lab and made measurements, whether they are mass, volume, or length, how 
do you know if they are correct? Accuracy is the difference between a measured value and the accepted 
- or what we call the correct - value for that quantity. To improve accuracy, scientists will repeat the 
measurement as many times as is possible. Precision is a measure of how close all of these measurements 
are to each other. Therefore, measurements can have precision but not very close accuracy. An example of 
accuracy of measurements is having the following data: 26 mL, 26.1 mL, and 25.9 mL when the accepted 
value is 26.0 mL. This data also shows precision. However, if the data had been 25.2 mL, 25.0 mL, and 
25.2 mL, they would show precision without accuracy. 

Sample question: Jack collected the following volumes when doing a titration experiment: 34.25 mL, 
34.30 mL, 34.60 mL, 34.00 mL, and 34.50 mL. The actual volume for the titration required to neutralize 
the acid was 34.50 mL. Would you say that Jack's data was accurate? Precise? Both accurate and precise? 
Neither accurate nor precise? Explain. 

Solution: All of Jack's data would be accurate because they are close to the true value of 34.50 mL. The 
data would also be precise having only 2% variance between the highest number and the lowest number. 

The distinction between accuracy and precision and its importance in science is demonstrated in an An- 
nenberg video at Video on Demand - The World of Chemistry - Measurement (http : //www. learner . 
org/resources/series61 .html?pop=yes&pid=793#) 

Lesson Summary 

• The task in the experiment determines the unit of measurement; this then determines the piece of 
equipment. Example: If mass is to be measured, a balance will be chosen as the piece of equipment. 

• Conversely, the piece of equipment chosen will determine the unit of measurement. Example: If a 
graduated cylinder is chosen, the unit of measurement will be volume (mL or L). 

• Each piece of equipment has a specified number of significant digits to which it is able to measure. 
Example: A household thermometer may measure to ± VC or ± 1°F, where as an ordinary high 
school alcohol thermometer measures to ± 0.1°C. 

• Significant digits are used in all parts of quantitative measurements in science. Five main rules are 
provided to read the significant digits of numbers and two main rules for solving algebraic equations 
maintaining proper significant digits. 

www.ckl2.org 82 



• Accuracy is how close the value is to the actual value (remember A and a). 

• Precision is how close values are in an experiment to each other. Precision is dependent on the 
significant digits of the instrument or measurement. 

Review Questions 

1. Suppose you want to hit the center of this circle with a paint ball gun. Which of the following are 
considered accurate? Precise? Both? Neither? 




2. Four students take measurements to determine the volume of a cube. Their results are 15.32 cm 3 , 
15.33 cm 3 , 15.33 cm 3 , and 15.31 cm 3 . The actual volume of the cube is 16.12 cm 3 . What statement(s) 
can you make about the accuracy and precision in their measurements? 

3. Find the value of each of the following to the correct number of significant digits. 



(a) 


1.25 + 11 


(b) 


2.308 


-1.9 


(c) 


498- 


97.6 


(d) 


101.3 


-=-12 


(e) 


25.69 


xO.51 



4. Why is the metric system used in chemistry? 

5. Distinguish between accuracy and precision. 

6. How many significant digits are present in each of the following numbers: 

(a) 0.002340 

(b) 2.0X10- 2 

(c) 8.3190 

(d) 3.00 x10 s 

7. Nisi was asked the following question on her lab exam. When doing an experiment, what term best 
describes the reproducibility in your results? What should she answer? 

(a) accuracy 

(b) care 

(c) precision 

(d) significance 

(e) uncertainty 

83 www.ckl2.org 



8. Karen was working in the lab doing reactions involving mass. She needed to weigh our 1.50 g of each 
reactant and put them together in her flask. She recorded her data in her data table and began to 
look at it (Table 3.4). What can you conclude by looking at Karen's data? 

(a) The data is accurate but not precise. 

(b) The data is precise but not accurate. 

(c) The data is neither precise nor accurate. 

(d) The data is precise and accurate. 

(e) You really need to see the balance Karen used. 

Table 3.4: 

Mass of Reactant 1 Mass of Reactant 2 

Trial 1 1.45 + 0.02 g 1.46 + 0.02 g 

Trial 2 1.43 ± 0.02 g 1.46 + 0.02 g 

Trial 3 1.46 + 0.02 g 1.50 ± 0.02 g 



9. Find the value of each of the following to the correct number of significant digits. 

(a) 3.567 + 3.45 

(b) 298.968 + 101.03 

(c) 1.25x11 

(d) 27^5.67 

(e) 423x0.1 

Further Reading / Supplemental Links 

The website listed below offers lessons, worksheets, and quizzes on many topics in high school chemistry. 

• Lesson 2-2 is on Accuracy and Precision, http : //www . f ordhamprep . org/gcurran/sho/sho/lessons/ 
lessonl2.htm 

• http : //learner . org/resources/series61 . html 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. The website has two 
videos that apply to this lesson. One is a video called Measurement, The Foundation of Chemistry 
that details the value of accuracy and precision. Another video called Modeling the Unseen relates to 
the scientific method and the use of models. 

Vocabulary 

significant digits A way to describe the accuracy or precision of an instrument or measurement. 

accuracy How close a number is to the actual or predicted value. 

precision How close values are in an experiment to each other. 
www.ckl2.org 84 



Labs and Demonstrations for Making Measurements 
Teacher's Pages for Thermometer Calibration 

Investigation and Experimentation Objectives 




In this activity, the student will determine one case of experimental error and recognize the limitations of 
measuring instruments. 

Notes: 

Thermometers should be stored vertically when not in use. You can stand them upright in a large beaker 
or in tall test tube racks. When thermometers are stored horizontally, they sometimes suffer a separation 
of the liquid near the top. If you have thermometers with separated liquid, you can sometimes shake them 
down or bounce the bulb gently on a folded towel to rejoin the liquid. 

Answers to Pre-Lab Questions 

1. Why is a mixture of ice and water, rather than ice alone, used in calibrating a thermometer? 

You cannot be sure of the temperature of solid ice. It might be — 10°C When ice and water are both 
present and in equilibrium, you can be sure the temperature of the mixture is 0°C. 

2. Why does the boiling point of a liquid vary with the barometric pressure? 

Water boils when its vapor pressure is equal to the surrounding pressure. If the surrounding pressure is 
above or below normal atmospheric pressure, then the boiling point of a liquid will be above or below its 
normal boiling point. 

3. What is the approximate boiling point of pure water at 380 Torr? 
Between 80° C and 82° C. 

4. What is the approximate boiling point of pure water at 800 Torr? 
Near 101°C 

5. Food products such as cake mixes often list special directions for cooking the products in high altitude 
areas. Why are special directions needed? Would a food product needing such directions require a longer 
or shorter time period to cook under such conditions? 

At high altitudes, the atmospheric pressure is less than normal atmospheric pressure and therefore, the 
boiling point of water is below 100°C Since boiling water is less than 100°C, cooking in boiling water will 
take longer. 

Table 3.5: Vapor Pressure of Water at Various Temperatures 

Temperature in °C Vapor Pressure in mm Temperature in °C Vapor Pressure in mm 

of Hg of Hg 

-10 2.1 52 102.1 

-5 3.2 54 112.5 

4.6 56 126.8 

2 5.3 58 136.1 

4 6.1 60 149.4 

6 7.0 62 163.8 

8 8.0 64 179.3 

85 www.ckl2.org 



Table 3.5: (continued) 



Temperature in °C 



Vapor Pressure in mm 
of Hg 



Temperature in °C 



Vapor Pressure in mm 
of Hg 



10 

12 
14 
16 

18 
20 
22 
24 
26 
28 
30 
32 
34 
36 
38 
40 
42 
44 
46 
48 
50 



9.2 

10.5 

12.0 

13.6 

15.5 

17.5 

19.8 

22.4 

25.2 

28.3 

31.8 

35.7 

39.9 

44.6 

49.7 

55.3 

61.5 

68.3 

75.7 

83.7 

92.5 



66 

68 

70 

72 

74 

76 

78 

80 

82 

84 

86 

88 

90 

92 

94 

96 

98 

100 

102 

104 

106 



196.1 
214.2 
233.7 
254.6 
277.2 
301.4 
327.3 
355.1 
384.9 
416.8 
450.9 
487.1 
525.8 
567.0 
610.9 
657.6 
707.3 
760.0 
815.9 
875.1 
937.9 



Thermometer Calibration 

Background Information 

The most common device for measuring temperature is the thermometer. The typical thermometer used in 
general chemistry labs has a range from -20°C to 120°C Most laboratory thermometers are constructed of 
glass and therefore are very fragile. Older thermometers contain mercury as the temperature sensing liquid 
while newer thermometers contain a red colored fluid. The mercury thermometers are hazardous if they 
break because mercury vapors are poisonous over long periods of inhalation and the mercury vaporizes 
slowly and so when it is spilled, the lab is toxic for several months unless every drop of mercury is picked 
up. The red colored liquid thermometers are also hazardous if they break because the liquid is flammable 
and may be toxic. Great care should be exercised when handling thermometers of either kind. 

The typical laboratory thermometer contains a bulb (reservoir) of temperature sensing fluid at the bottom; 
it is this portion of the thermometer which actually senses the temperature. The glass barrel of the 
thermometer above the liquid bulb contains a fine capillary opening in its center, into which the liquid 
rises as it expands in volume when heated. The capillary tube in the barrel is very uniform in its cross- 
section all along the length of the thermometer. This insures that the fluid will rise and fall uniformly 
when heated or cooled. 

(NOTE: laboratory thermometers look like clinical thermometers for taking people's temperatures but they 
are not the same. The clinical thermometer has a constriction in the tube so that after the temperature goes 
up and the thermometer is removed from the heat source, the liquid will not go back down. Such clinical 
thermometers must be shaken to lower the temperature reading before each use. Lab thermometers have 



www.ckl2.org 



86 



no such constriction and hence the temperature reading immediately starts down when the heat source 
is removed. For that reason, lab thermometers must be read while the bulb is still in contact with the 
material whose temperature is being taken.) 

Because thermometers are so fragile, it is a good idea to check them, now and then, to make sure they 
are still working properly. To check a thermometer, a process of calibration is used. To do this, you 
will determine the reading given by your thermometer in two systems whose temperature is known with 
certainty. If the readings of your thermometer differ by more than one degree from the true temperatures, 
it should be removed from use. 

A mixture of ice and water which has reached equilibrium has a temperature of exactly 0°C and will be used 
as the first calibration point. The second calibration point will be boiling water whose exact temperature 
must be determined using the barometric pressure in the lab. 

Pre-Lab Questions 

1. Why is a mixture of ice and water, rather than ice alone, used in calibrating a thermometer? 

2. Why does the boiling point of a liquid vary with the barometric pressure? 

3. What is the boiling point of pure water at 380 Torr? 

4. What is the boiling point of pure water at 800 Torr? 

5. Food products such as cake mixes often list special directions for cooking the products in high altitude 
areas. Why are special directions needed? Would a food product needing such directions require a 
longer or shorter time period to cook under such conditions? 

Purpose 

In this experiment, you will check a thermometer for errors by determining the temperature of two stable 
equilibrium systems. 

Apparatus and Materials 

• Thermometer 

• 400 mL beaker 

• 250 mL beaker 

• distilled water 

• ice 

• hot plate 

• stirring rod 

• boiling chips. 

Safety Issues 

Mercury thermometers are hazardous if they break because mercury vapors are poisonous over long periods 
of inhalation and the mercury vaporizes slowly and so when it is spilled, the lab is toxic for several months 
unless every drop of mercury is picked up. The red colored liquid thermometers are also hazardous if they 
break because the liquid is flammable and may be toxic. Great care should be exercised when handling 
thermometers of either kind. 

Procedure 

Fill a 400 mL beaker with ice and add tap water until the ice is covered with water. Stir the mixture is a 
stirring rod for one minute. Dip the thermometer into the ice water mixture so that the thermometer bulb 
is approximately centered in the mixture (not near the bottom or sides). Leave the thermometer in the 
mixture for two minutes and then read the thermometer to the nearest 0.2 degree while the thermometer 
is still in the ice water bath. Record the temperature. 

87 www.ckl2.org 



Allow the thermometer to return to room temperature by resting it is a safe place on the laboratory table. 

Half fill a 250 mL beaker with distilled water and place it on a hot plate. Add 2 or 3 boiling chips 
to the water. Heat the water to boiling. Dip the thermometer into the boiling water making sure the 
thermometer does not get near the bottom, sides, or top of the water. Hold it there for 2 minutes and 
record the temperature reading to the nearest 0.2 degree. 

Ask your instructor for the current barometric pressure reading in the laboratory room, look up the actual 
boiling point of water at this pressure and record. 

Data 

Actual freezing point of water = 

Freezing point determined by your thermometer = 

Difference between correct and trial values = 

Barometric pressure in the room = 



Actual boiling point of water at this pressure = 

Boiling point determined by your thermometer = 

Difference between correct and trial values = 

Post-Lab Questions 

1. Calculate the percent error of your measurement of the freezing point of water. 

actual value — trial value 



'A 



b error 



actual value 



xlOO 



2. Calculate the percent error of your measurement of the boiling point of water. 

_. actual value - trial value 

% error = X 100 = 

actual value 



Table 3.6: Actual Boiling Point of Water versus Various Room Pressures 



Room Pressure (mm of Hg) 



Boiling Point of Water (°C) 



750 
751 
752 
753 
754 
755 
756 
757 
758 
759 
760 
761 
762 
763 
764 
765 
766 

www.ckl2.org 



99.6 

99.7 

99.7 

99.8 

99.8 

99.8 

99.9 

99.9 

99.9 

100.0 

100.0 

100.1 

100.1 

100.1 

100.2 

100.2 

100.2 



88 



Table 3.6: (continued) 



Room Pressure (mm of Hg) Boiling Point of Water (°C) 

767 100.2 

768 100.3 

769 100.3 

770 100.3 



3.3 Using Data 

Lesson Objectives 



• Recognize patterns in data from a table of values, pictures, charts and graphs. 

• Make calculations using formulae for slope and other formulae from prior knowledge. 

• Construct graphs for straight lines. 

• Construct graphs for curves. 

• Read graphs using the slope of the line or the tangent of the line. 

Introduction 

Earlier, we learned about qualitative and quantitative observations, and that with quantitative observa- 
tions, we need measurements. In science, measurements mean data. In this upcoming section, we will 
delve deeper with data to look at patterns and to graph data. Sometimes we can graph the data, make 
calculations, sketch the line, and calculate the slope. All of these quantitative observations help us to 
formulate a conclusion that will be based on evidence. 

Recognizing Patterns in the Data 

As stated earlier, data can provide enormous information to scientists for making interpretations and 
drawing conclusions. In order for scientists to do this, they have to be able to look at a set of data 
and recognize patterns. Data can be in the form of pictures, charts, or graphs. Take for example, the 
picture found in (Figure 3.1). This, although not particularly chemistry related, has much to do with the 
concept of pattern recognition and the data gathered from these patterns. What do you believe scientists 
determined when first viewing this image prior to August 23, 2005? 

Now, let us look at a common chemistry example. We all know that metals are supposed to be chemically 
reactive. How do we know that? Well, that is a property of metals. But how reactive are they? How 
can you tell? What do you know about the periodic table that would help you determine this right now? 
Look at Table 3.7 and see if you can gather a little evidence that can solidify your conclusions about the 
chemical reactivity of metals and the periodic table. 

Table 3.7: Information about Metals and Reactivity 

Metal Chemical Reactivity in Atmosphere 

Sodium Stored in toluene, extremely reactive 

Potassium Burns in 02 in seconds 

Calcium Slower to react with Oi than its neighbors to the 

right 

89 www.ckl2.org 



Table 3.7: (continued) 



Metal 



Chemical Reactivity in Atmosphere 



Titanium 

Aluminum 

Gold 

Platinum 

Copper 

Iron 



Resists corrosion after forming an oxide barrier 

Resists corrosion after forming an oxide barrier 

Does not react with oxygen 

Does not react with oxygen 

Does not react with H2O, will react slowly with O2 

Rusts in 02 



What kinds of conclusions can you make from reading this table? Can you determine that the reactions 
between the metals and the oxygen in the air decrease going across a row on the periodic table? Did you 
notice that the alkali (group one) metals (sodium and potassium) are the most reactive of all the metals 
with oxygen and the alkaline earth (group two) metals (calcium) are the second most reactive? Yes, they 
are. Aluminum as well as titanium will actually become coated with an oxide of their metals, which acts 
like a protective shield against further reaction. Look at the figure below. The copper has turned green as 
a result of this protective coating that formed on the copper rooftop. Have you ever seen this before? 

The current Hotel Vancouver took over a decade to build during the 1930s as the Great Depression 
put a temporary halt to construction. It was Vancouver's tallest building from 1939 to 1972. (Source: 
http : //commons . wikimedia . org/wiki/File : HotelVancouver . jpg . CC-BY-SA) 




What kind of observations are these? Are they quantitative? No, of course not because they have no 
measurements attached; they are qualitative observations. 

What about other types of data that involve quantitative observations? Can we look at measurements and 
determine patterns? Yes we can. Statisticians, weather persons, stock market workers, sports analysts, 
and chemists (to name a few occupations) do this on a daily basis for their regular jobs. What if you were 
trying to identify an unknown substance based on a volume displacement experiment. You were given a 
series of known substances of known masses, and you then determined how much volume they displaced 
in a cylinder of water. Through your experiment, the following data was recorded (Table 3.8). 



www.ckl2.org 



90 




Figure 3.1: Hurricane Katrina taken on Aug. 28, 2005, at 11:45 a.m. EDT by NOAA when the storm was 
a Category Five hurricane. 

Table 3.8: Volume Displacement Data 



Substance 



Mass 



Volume ol Water Displaced 



Aluminum 
Iron 
Copper 
Silver 
Zinc 
Lead 



2.7 g 

7.86 g 
8.92 g 
10.5 g 
7.14 g 
11.34 g 



1.0 mL 

1.0 mL 
1.0 mL 
1.0 mL 
1.0 mL 
1.0 mL 



First, which metal do you think is the densest, just looking at the data table? Do you think it would be 
lead or aluminum? How can you tell? Lead is the right answer because there is a heavier mass of lead 
displacing the same volume of water. Did you notice that copper is heavier than iron, and iron is heavier 
than zinc? Now what kind of data do you think this is? Is it qualitative or quantitative? Quantitative is 
the right answer because we are dealing with measurements. 

An interesting note is that 19.3 g of gold would have displaced the same volume of water. Would it be the 
densest? Yes it would. It also means that this same amount of gold converts to approximately 212.5 lb/gal. 
Remember seeing in television or movies where the villain is running off with a bag full of gold bars? This 
would be rather difficult knowing the density of the metal from this table. Being able to read tables and 
data gives us the power to understand the world around us. 

Another way to generalize trends in data and to make interpretations based on these trends is to plot a 
graph. Graphs are like numerical pictures that provide an image of the data collected in an experiment. 
Suppose you were asked to record the temperature of a mixture as it was slowly heated in a hot water 
bath. You record the following data as you watch your experiment (Table 3.9). You then plot the data 
on a graph to see what the numbers tell you. 

Table 3.9: Time vs. Temperature 



Time (s) 



Temperature (°C) 




1 
2 
3 

4 

5 
6 

7 
8 



23.5 
24 
25 
26 
27 
9128 
29 
30 
31 



www.ckl2.org 




Temperature 

CO 



-I — h 



Time (s) 



1 1 

Looking at the graph, what do you notice about the temperature of the mixture as it is heated in the hot 
water bath? It is constantly increasing. What does the initial point, or the y— intercept represent? It is the 
value of the initial temperature, most likely room temperature. Using a graph, it is clear to see that the 
y—axis, in this case temperature, is dependent on the x-axis. For our example, the independent variable 
is time. What does this mean? It means that for a change to be made in temperature, time must pass. 

Being able to plot tables of values and read the corresponding graphs is an important skill, not only for 
mathematics but also for science. Interpretations can be made by using either of these representations; one 
may be more visual and thus sometimes easier to interpret than the other. Now try one example where 
you have to plot a graph and make some interpretations. 

Sample Question: The following data represents the marks of 12 students in a Math Test and in a Chemistry 
Test. All Marks are out of 50. 

Math Marks vs. Chemistry Marks 



Math 


17 


38 


40 


17 


28 


30 


45 


24 


48 


42 


32 


36 


Chemistry 


8 


32 


36 


17 


19 


20 


43 


16 


48 


40 


22 


29 



a. Plot this data on an x - y axis, with Math marks on the x-axis. 

b. Draw the line of best fit. A line of best fit is drawn on a scatter plot so that it joins as many points 
as possible and shows the general direction of the data. When constructing the line of best fit, it is also 
important to keep, approximately, an equal number of points above and below the line. 

c. Estimate the chemistry mark of a student who scored 32 on a math test. 

d. Estimate the math mark of a student who scored 45 on a chemistry test. 

e. Based on the trend found in the data, what can you say about the relationship between math and 
chemistry marks? 

Solution: a) and b) 

Graph 2: Math Marks vs. Chemistry Marks 



www.ckl2.org 



92 




c) The chemistry mark when a student makes 32 in math is approximately 25. We can see this by the 
blue line. By interpolating the data, we can draw a line up from the 32 mark on the x-axis to where the 
line of best fit runs through the data points. Here, we cut across to the y-axis to find the corresponding 
chemistry mark. 

d) The math mark when a student makes 45 in chemistry is approximately 47. We can see this by the 
green line. By interpolating the data, we can draw a line across from the 45 mark on the y-axis to where 
the line of best fit runs through the data points. Here, we draw a line down to the x-axis to find the 
corresponding math mark. 

e) The trend shows that as math marks increase, so do chemistry marks. The dependent variable in this 
graph is the chemistry mark. 



Making Calculations With Data 

Being able to recognize patterns from tables, charts, pictures, and graphs is a worthy skill for any scientist 
and science student. By having charts, pictures, tables, and graphs, you can also perform a large variety 
of calculations depending on the independent and dependent variables. From here we can accomplish such 
things as making further predictions, drawing more conclusions, or identifying unknowns. Let's say, for 
example, in the density experiment from earlier, we were given an unknown for our experiment. Now, take 
the same chart but add a fourth column representing the density of the metal (Table 3.10). Recall the 
formula for density: 



Density 



mass 
volume 



Now, filling in the fourth column, using the density formula for our experimental data, we see the following 
information. Remember density = mass/volume or, in the table below, column 4 = column 2/ column 3. 

Table 3.10: 



Substance 



Mass 



Volume of Water Dis- Density 
placed 



Aluminum 



2.7 g 



1.0 mL 



2.7 g/mL 



93 



www.ckl2.org 



Table 3.10: (continued) 



Substance Mass Volume of Water Dis- Density 

placed 

Iron 7.86 g 1.0 mL 7.86 g/mL 

Copper 8.92 g 1.0 mL 8.92 g/mL 

Silver 10.5 g 1.0 mL 10.5 g/mL 

Zinc 7.14 g 1.0 mL 7.14 g/mL 

Lead 11.34 g 1.0 mL 11.34 g/mL 

Unknown 1.78 g 0.2 mL 8.9 g/mL 

Just by doing this calculation, we can identify our unknown in the experiment? Sure, it is copper. Calcula- 
tions are frequent in chemistry, as you will learn. Most times you will see we will use a variety of formulae 
to solve problems similar to those you would solve in any course. The relationship from comparing sets of 
answers give us the interesting parts of these types of calculations. 

Sample Question: How long does it take you to run 3.5 miles at 7 mph? 

Solution: We know that: speed = di t s ^ ce 

Therefore: 

distance 
time 



time 



speed 
3.5 mi 



7 mi/h 
time = 0.5 h 

Another important skill with calculations is converting from one unit to another. Frequently in chemistry, 
data will appear in problems that require us to use conversion factors before completing the problem. Some 
conversion factors include 100 cm = 1 m, 1000 mL = 1 L, and 1 km = 1000 m. 

Sample Question: The speed of light is 3.00 X 10 8 m/s. The speed of sound is 1230 km/h. How much 
faster is the speed of light than the speed of sound. 

Solution: Speed of Sound = 1230 km/h x ±f2Lm x _L^_ x |>5i 

Speed of Sound = 342 m/s 

speed of light _ 3.00 x 10 8 m/s 

speed of sound 342m/s 

speed of light _ ^ 

B - 8.77 x 10° 



speed of sound 



In other words, the speed of light is almost 900, 000 times faster than the speed of sound. Amazing isn't 
it! 

Preparing Graphs From Data 

Most times in the laboratory, we collect data of some sort and then carry it back to our desk to analyze. We 
want to determine the melting point of an unknown solid, so we take the melting points of various knowns 
and then that of our unknown; following this, we make a table, finally writing the data gathered from our 

www.ckl2.org 94 



experiment into the table. Some laboratory experiments that we, as scientists, do require us to draw graphs 
in order to interpret the results and make any conclusions. Drawing a graph that anyone can understand 
is a useful skill to any scientist. Graphs have to be properly labeled on the x-axis (the horizontal) and the 
y-axis (the vertical). The graphs should indicate a straight line or smooth curve indicating that the data 
is continuous. A straight line represents a linear relationship; a curved line does not. 

Sample Question: Medical practitioners have been studying the heart for a long time. As a result, we can 
now calculate your heart rate based on a formula derived from your age. Plot the table of values given 
below. Properly label the graph include the independent and dependent variable. Draw the line all the 
way to the y-axis so that you can find the y- intercept. Finally, find your age on the x-axis and then find 
your maximum heart rate by drawing a vertical line up to the graph. 

Age in Years (x) 30 40 50 60 

Beats Per Minute (y) (maximum heart rate) 90 80 70 60 

Solution: At the age of 17, the maximum heart rate (beats per minute) read from the graph is 103. 
Heart Rate vs Age 



Maximum Heart Rate 
(beats per minute) 
Dependent 




Age (Years) 
Independent 



Reading Results From the Graph 

Many properties in chemistry lead to linear relationships when plotted. We saw this with the temper- 
ature/time relationship. Other properties in chemistry do not form this linear relationship. Take for 
example the relationship between concentration and temperature. Remember the last time you made a 
cup of instant coffee or hot chocolate? Why did you boil the water? What would have happened if you used 
warm tap water or even cold water from the refrigerator? Putting aside the anticipated taste difference, 
what would have happened to the solid you were trying to dissolve? The amount of instant coffee or hot 
chocolate powder in your cup that actually dissolves in warm or cold water would be small compared to 
when you use boiling water. This property is known as solubility. The solubility of a substance is the 
amount that can dissolve in a given amount of solution. Solubility is affected by the temperature but rarely 
linearly. Look at the data table below (Table 3.11). This data is for the solubility of KCIO3 (potassium 
chlorate) in water. 



95 



www.cki2.0rg 



Table 3.11: 



Temperature (°C) 



Solubility ( g/100 mL H 2 0) 





20 

40 

60 

80 

100 



3.3 

7.3 

13.9 

23.8 

37.5 

56.3 



Now graph the data and see what kind of curve we get. 
Solubility of Potassium Chlorate 




Notice how when the line is drawn, the relationship between grams of potassium chlorate that dissolve 
in 100 mL of water and temperature is not linear but curved. We can still interpret the data as we did 
earlier. For example, what is the solubility of KCIO3 (how much KCIO3) when the temperature is 75°C or 
at room temperature. 



Solubility of Potassium Chlorate 



www.ckl2.org 



96 




Using the same procedure, we draw a line up from 75°C and then over to the v-axis. It reads 34.5 g/100 mL. 
Therefore, 34.5 g of KCIO3 can dissolve in 100 mL of H2O at 75° C. 

Now you try one: 

Sample Question: 

Ammonia, NH3, and sodium chloride, NaCl, are known to have the following solubility data (Table 3.12). 

Table 3.12: 



Temperature (°C) 



NH 3 Solubility ( g/100 mL H 2 0) NaCl Solubility ( g/100 mL H 2 0) 





20 

40 

60 

80 

100 



88.5 
56.0 
34.0 
20.0 
11.0 
7.0 



35.7 
35.9 
36.4 
37.1 
38.0 
39.2 



a) Properly graph the data for each substance. 

b) Are either of these linear? Explain. 

c) What would be the solubility of each of the substances at 50°C? 
Solution: 

Solubility of NH 3 and NaCl 

97 



www.cki2.0rg 



70 
60 

Solubility 
in water 50 
(g/mL) 













































N'S- 
























































































































































































Na 


:i 








i 


1 





























































































































































S 10 15 20 25 30 35 40 45 50 55 60 65 70 75 BO 65 90 95 100 

Temperature ( C) 



b) The Solubility data for NaCl represents a linear relationship when graphed. You can see this with the 
blue line in the graph above. The NH% line (in red) is curved; therefore it is non-linear. 

c) The solubility of NH 3 at 50°C is 27 g/100 mL H 2 0. The solubility for NaCl at 50°C is 37.5 g/100 niL H 2 0. 
The graph below is marked to show the line travelling up from 50°C and over to the y-axis (solubility) to 
find the answer for both of these parts. 

Solubility of NH 3 and NaCl 



70 
60 

Solubility 
in water 50 
(g/mL) 

37.5 
30 





































































































































































































































Na 


:i 







































































"--■-, 



























































































































5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 

Temperature { C) 



30 B5 90 95 100 



We can do more than just graph and read graphs of linear and non-linear data in order to make conclusions. 
For this, we use formulas like the slope of a line. Remember slope from math class? It is a formula used 
to find the rate at which one factor is affecting the other, either positively or negatively. Remember the 
formula for slope from math class? 



rise y 2 - yi 

slope = or m = 

run X2 - x\ 



Let's look at how the slope formula can be used on a graph to see how one factor is affecting another in 
an experiment. 

A modern hot air balloon. (Source:http : //en . wikipedia . org/wiki/ Image : 2006_0 j iya_balloon_f est ival_ 
011.jpg- CC-BY-SA) 



www.ckl2.org 



98 




In the 18005, the use of the hot air balloons was extremely popular as a sport as well as an extracurricular 
activity for those who could afford the luxury. Up to this point, the study of the relationship of gases and 
the factors of temperature, pressure, and volume was limited to Robert Boyle's experiments with pressure 
and volume. Jacques Charles came along with his experiments on the relationship between volume and 
temperature. Here is some typical data from a volume/temperature experiment with gases (Table 3.13). 

Table 3.13: 



Temperature (C°0) 



Volume of Gas (cm 3 



20 

40 

60 

80 

100 

120 



60 
65 

70 

75 
80 
85 



Volume of a Gas vs. Temperature 

I I I I I I II 




99 



www.ckl2.org 



Now, find out what affect the temperature has on the volume of the gas. In other words, find the slope. 
Pick two points that are on the line and use the equation above to find the value of m. 



y2- 


yi 




X2 - 


XI 




80- 


-65 




100 


-40 




15 






" 60 






= 0.25 


cm 3 /' 


C 



What does this mean? It means that for each increase in temperature of 1°C, the volume increased by 
0.25 cm 3 (or 0.25 mL). This translates to approximately 1 mL increase every 4°C. This is a positive 
increase (notice the slope is increasing or going up). Now let's try another one. 

Look at Table 3.14 for a set of data from an experiment performed between bromine and formic acid in 
a laboratory setting. The reaction was performed to see if the decrease in bromine concentration could 
cause the reaction to subsequently slow down. In other words, if they took some of the bromine out of the 
reaction, would the reaction start to slow down? Look at the data and see what happened. 

Table 3.14: 

Reading Concentration of Bromine Time 

(mol/L) 

1 0.1 

2 0.07 0.75 

3 0.05 1.75 

4 0.035 2.49 

5 0.02 3.48 

6 0.01 5 

7 0.005 6.2 

8 0.001 7.5 

9 0.0 8.8 

10 0.0 9 



Look at the data table, can you tell if an increase in bromine concentration had an effect on the rate of a 
reaction? Did it make it go faster or slower. Let's take a look. From 7\ to T-j the bromine concentration 
decreased from 0.1 mol/L to 0.005 mol/L, a decrease of 0.095 mol/L. The time it took for this decrease 
was 6.2 seconds. What does this tell us about the rate? A preliminary conclusion would be that a decrease 
in bromine concentration causes the rate of the reaction to also decrease. A graph might make it a little 
easier to make conclusions based on the data. 

Concentration of Bromine vs. Time 
www.ckl2.org 100 




Using slope to calculate the effect here is not as easy because we have a curve. We can easily see that as 
the concentration increases, so does the rate, but by how much? The slope would actually tell us that. In 
order to determine the slope of a curve, you need to draw in a tangent to the curve. A tangent is just a 
straight line drawn to the curve; from this you would calculate the slope. Look at the graph below, the 
tangents are drawn in using a red pen. 

Concentration of Bromine vs. Time 




If we draw a tangent, look, we have a straight line. We can now, find two points. 



m 



m 



y2-yi 

x 2 -xi 
0.06 - 0.07 



1.1-0.75 
m = -0.028 mol/L • s 



This means that as the concentration of bromine decreases, so does the rate of the reaction. Look at the 
units for the slope. The units are mol/L s. These are the units for rate. This means, interestingly enough, 
that as the concentration goes down, the reaction slows down. 



101 



www.ckl2.org 



Lesson Summary 

• Patterns can be found in data sets, pictures, charts, and graphs. From here scientists can make 
interpretations of the data and draw conclusions. 

• When drawing graphs of tables of values, a straight line or a smooth curves can be drawn. Some 
data sets do require a line of best fit. 

• A line of best fit is drawn on a scatter plot so that it joins as many points as possible and shows the 
general direction of the data. When constructing the line of best fit, it is also important to keep, 
approximately, an equal number of points above and below the line. 

• Conversion factors are necessary for calculations where the units do not match. For example, km 
and m. 

• Recall the slope formula: m = } ' 2 yi 

• For curved lines, remember to draw the tangent first and then find the slope of the tangent line. 

Review Questions 

1. Why is the slope of a graph so important to chemistry? 

2. What would you do to find the slope of a curved line? 

3. What is a conversion factor used for? 

4. Of the following professions, choose the one that uses data to find the identity of unknown finger- 
prints? 

(a) analytical chemist 

(b) archaeological chemist 

(c) inorganic scientist 

(d) forensic scientist 

(e) quality control chemist 

5. Which speed is the slowest? 

(a) 200 m/min (200 m/min X 1 km/1000 m = 0.2 km/min) 

(b) 0.2 km/min 

(c) 10 km/h (10 km/h X 1 h/6 min = 0.17 km/min) 

(d) 1.0 X 10 5 mm/min (1.0 X 10 5 mm/min X 1 m/1000 mm X 1 km/1000 m = 0.1 km/min) 

(e) 10 mi/h (10 mi/h X 1.603 km/mi X lh/60 min = 0.267 km/min or 0.27 km/min) 

6. Andrew was completing his density lab for his chemistry lab exam. He collected the following data 
in his data table (Table 3.15). 

(a) Draw a graph to represent the data. 

(b) Calculate the slope. 

(c) What does the slope of the line represent? 

(d) Can you help Andrew determine what his unknown is by looking in a standards table? 



www.ckl2.org 102 



Table 3.15: 



Mass of Solid (g) Volume of Solution (niL) 

3.4 0.3 

6.8 0.6 

10.2 0.9 

21.55 1.9 

32.89 2.9 

44.23 3.9 

55.57 4.9 



7. Donna is completing the last step in her experiment to find the effect of the concentration of ammonia 
on the reaction. She has collected the following data from her time trials and is ready for the analysis 
(Table 3.16). Donna is now required to graph the data, describe the relationship, find the slope and 
then discuss the meaning of the slope. Help Donna with the interpretation of her data. 

Table 3.16: 

Time (s) Concentration (mol/L) 

0.20 49.92 

0.40 39.80 

0.60 29.67 

0.81 20.43 

1.08 14.39 

1.30 10.84 

1.53 5.86 

2.00 1.95 

2.21 1.07 

2.40 0.71 

2.60 0.71 

8. 

Further Reading / Supplemental Links 

• http : //en . wikipedia . org/wiki/Metal 

• http : //en . wikipedia . org/wiki/Hurricane_Katrina 

• http : //en . wikipedia . org/wiki/Solubility_table 

Vocabulary 

chemical reactivity An observation of the behavior of the element of compound based on its position 
in a reactivity (or activity) series. 

periodic table An arrangement of elements in order of increasing atomic number. 

103 www.ckl2.org 



alkali metals Group 1 metals of the periodic table (H,Li,Na,K,Rb,Cs,Fr). 

alkaline earth metals Group 2 metals of the periodic table (Be,Mg,Ca,Sr,Ba,Ra). 

density Measurement of a mass per unit volume. Density = rpr^- . 

graphs Pictorial representation of patterns using a coordinate system (x — y axis). 

dependent variable The variable that changes depending on another variable (y-axis variable). 

independent variable The variable that changes to cause another variable to change (x-axis variable). 

y-intercept Where the line crosses the y-axis. 

conversion factor A ratio used to convert one unit to another. 

linear relationship A relationship where the x-values change proportionally with the y-values leading 
to a straight line. 

non-Linear relationship A relationship where the x-values do not change proportionally with the y- 
values leading to a curved line. 

a line of best fit A line drawn on a scatter plot so that it joins as many points as possible and shows 
the general direction of the data. When constructing the line of best fit, it is also important to keep, 
approximately, an equal number of points above and below the line. 

slope A formula to find the rate at which one factor is affecting the other. 

y2-vi 
m = 

X2 - Xl 

tangent A straight line drawn to the curve. 

solubility The amount of a substance that can dissolve in a given amount of solution. 

Labs and Demonstrations for Using Data 
Density Determination 

Investigation and Experimentation Objectives 




In this activity, the student will make and record observations using rulers, balances, and graduated 
cylinders. The student will use his measurements to mathematically calculate the density of both regular 
and irregular objects. 

Pre-Lab Discussion 

Density is defined as the mass per unit volume of a substance. The table below lists the densities of some 
well known substances. 

www.ckl2.org 104 



Table 3.17: Density of Some Common Substances 



Substance 



Density 



Substance 



Density 



Water 


1.0 g/cm 3 


Aluminum 


Oxygen gas 


0.0013 g/cm 3 


Iron 


Sugar 


1.6 g/cm 3 


Lead 


Table salt 


2.2 g/cm 3 


Gold 



2.7 g/cm 3 
8.9 g/cm 3 
11.3 g/cm 3 
19.3 g/cm 3 



Density measurements allow scientists to compare the masses of equal volumes of substances. If you had 
a piece of lead as large as your fist and a piece of gold as large as your thumb, you would not know which 
substance was innately heavier because the size of the pieces are different. Determining the density of the 
substances would allow you to compare the masses of the same volume of each substance. The process for 
finding the density of a substance involves measuring the volume and the mass of a sample of the substance 
and then calculating density using the following formula. 



Density 



mass in grams 
volume in mL 



Example: Calculate the density of a piece of lead whose mass is 226 grams and whose volume is 20.0 mL. 
Also calculate the density of a sample of gold whose mass is 57.9 grams and whose volume is 3.00 mL. 

Solution: 



Density of Lead 
Density of Gold 



mass 226 g 

volume 20.0 mL 

mass 57.9 g 

volume 3.00 mL 



11.3 g/mL 
19.3 g/mL 



Methods of Measuring Mass and Volume 

The mass of substances is measured with a balance. In the case of a solid object that will not react with 
the balance pan, the object may be placed directly on the balance pan. In the case of liquids or reactive 
solids, the substance must be placed in a container and the container placed on the balance pan. In order 
to determine the mass of the substance, the mass of the container is determined before hand (empty) and 
then the container's mass is subtracted from the total mass to determine the mass of the substance in the 
container. There are several common procedures for determining the volume of a substance. The volume 
of a liquid is determined by pouring the liquid into a graduated cylinder and reading the bottom of the 
meniscus. For a regularly shaped solid, the volume can be calculated from various linear measurements. 



CUBE 



SPHERE 



CYLINDER 



: 



V-OOWW 




CZZ2 

y= 7ir 2 h 



For an irregularly shaped object, a graduated cylinder is partially filled with water and the volume mea- 
sured. The object is then submerged in the water and the new volume measured. The difference between 
the two volumes is the volume of the submerged object. 



105 



www.ckl2.org 



Equipment: Specific gravity blocks, graduated cylinders (10 mL and 100 mL), thread, millimeter ruler, 
balance, distilled water, glycerol. (If you have an overflow can, it also works well for submersion.) 

Procedure: 

1. Obtain a regularly shaped object from your teacher. Measure and record its mass in grams and its 
dimensions in centimeters. 

2. Add approximately 50 mL of tap water to a 100 - mL graduate and record its exact volume. Tie a 
thread to the block and carefully immerse it in the cylinder of water. Record the new volume in the 
cylinder. 

3. Measure and record the mass of a clean, dry, 10 - mL graduated cylinder. 

4. Add exactly 10.0 mL of distilled water to the cylinder. Measure and record the combined mass of the 
cylinder and water. 

5. Repeat steps 3 and 4 with glycerol instead of water. 

Data Table 
Object 

code name or letters 

width 

height 

length 

volume of water before block 

volume of water after block 

Distilled Water 

volume 

mass of empty graduate 

combined mass 

Glycerol 
volume 



mass of empty graduate 

combined mass 

Calculations 

1. Find the volume of the solid object using the dimensions and appropriate formula. 

2. Find the volume of the solid object using water displacement. 

3. Find the density of the block using the volume from calculation 1. 

4. Find the density of the block using the volume from calculation 2. 

5. Find the density of the distilled water. 

6. Find the density of the glycerol. 

7. If your teacher gives you the accepted values for the densities in this lab, calculate the percent error 
for your values. 

(experimental value) — (actual value) 

% error = — X 100 

(actual value) 

www.ckl2.org 106 



Demonstration of Density of Diet Soda vs. Regular Soda 

Investigation and Experimentation Objectives 




In this activity, the student will make observations and use critical thinking to suggest a possible explanation 
for the observations. 

Brief description of demonstration 

A can of diet soda and regular soda are place into a clear container of water. The diet soda floats while 
the regular soda sinks. 

Materials 

• 12 oz. can of diet soda 

• 12 oz. can of regular soda, preferably the same brand 

• Clear container with enough volume so the can has room to sink totally 

Procedure 

Fill the clear container to within 5 cm of the top with water. Place the diet soda into the container. It will 
float. Place the regular soda into the container. It will sink. 

Hazards 

None. 

Disposal 

Pour the water down the sink. 

Discussion 

This is a good demonstration of density and for the discussion of dependent and independent variables. 
The only significant difference between the cans is their contents. One may want to try different sodas of 
both diet and regular varieties to show this. 

Absolute Zero Determination Demo 

Investigation and Experimentation Ojectives 



k 



In this activity, the student will make and record observations, graph the measurement data, and make 
use of the concepts of interpolation and extrapolation in the construction of a graph. 

Brief description of demonstration 

An Absolute Zero Apparatus is placed in various liquids at different temperatures. The temperatures of 
each solution are known. The pressure is read from the pressure gauge on the apparatus. A graph is 
made with Celsius temperature on the vertical axis and pressure on the horizontal axis. The plot is then 
extrapolated to zero pressure. The extrapolated line will cross the temperature axis at absolute zero. 

Apparatus and Materials 

• Absolute zero apparatus (available from science supply companies for around $150) 

107 www.ckl2.org 




Figure 3.2 



www.ckl2.org 



108 



3 - Pyrex or Kimex 2.0 liter beakers 

Hotplate 

Ice 

Dry ice (you can find dry ice suppliers on the internet - dry ice can be stored in a Styrofoam cooler 

but do not put the lid on tightly) 

Ethanol - 600 mL 

If you have a mercury thermometer that covers the range -100°C to +100°C, you can use it to 

measure the temperatures of the baths. 



Procedure 



1. Fill one of the Pyrex beakers half full of tap water and place it on the hotplate to boil. 

2. Fill the another Pyrex beaker half full of tap water and crushed ice. 

3. Fill the third beaker about one-fourth full will broken pieces of dry ice and then add ethanol slowly 
(lots of fog) until the beaker is about half full. 

4. Place the bulb of the Absolute Zero Apparatus into the boiling water and leave it there until a 
constant pressure is reached. Record the temperature of the bath (taken to be 100°C) and the 
pressure on the gauge. 

5. Place the bulb of the Absolute Zero Apparatus into the ice water and leave it there until a constant 
pressure is reached. Record the temperature of the bath (taken to be 0°C) and the pressure on the 
gauge. 

6. Place the bulb of the device into the dry ice and alcohol slush and leave it there until a constant 
pressure is reached. Record the temperature (known to be — 81°C) and the pressure on the gauge. 

7. Plot a graph of the temperatures and pressures recorded. Make sure that the temperature axis on 
your graph extends below — 280°C. 

8. After the three points are plotted on the graph, lay a straight edge on the graph line and extend it 
to the zero pressure line. You will get a graph similar to the one shown below. 

109 www.ckl2.org 



100 

so 

60 
40 

:o 
o 

-88 

-40 

o W _«. 
jf 

9 -80 

S 

g -100 

p. 

a im 

s 

H -140 
-1«0 
-180 

-200 
-220 
-240 
-2«0 
280 
-300 





























































































































































































































































































































































































































































# 














































1 
9 


' 














































* 
f 














































* 














































, 


t 














































$ 














































a 


1 














































i 














































r 
















































t 

































































































0.1 0.2 0.3 0.4 0.5 0,6 0.7 0.8 1.0 1.2 1.3 1,4 1.5 1.6 1,7 l.S 1,9 2.0 

Pressure, atm. 

Hazards 

Do not handle dry ice with bare hands. Pot holders or thermal protection gloves are necessary to handle 
the boiling water beaker and the dry ice - alcohol slush beaker. 

Disposal 

Once the dry ice has all melted, the solutions can be poured down the sink. 

Discussion 

Pressure is caused by the collisions of gas particles with each other and the walls of their container. When 
the temperature is lowered, the particles move more slowly, decreasing the frequency and strength of these 
collisions. In turn, the pressure falls. 

Absolute temperature can be defined as the temperature at which molecules cease to move. Therefore 
absolute zero temperature corresponds to zero pressure. 

Extending a graph beyond actual data points is called extrapolation ... a not-always acceptable procedure. 

3.4 How Scientists Use Data 



Lesson Objectives 



Define the terms law, hypothesis, and theory. 
Explain why scientists use models. 



www.ckl2.org 



110 



Introduction 

In the last section, we learned a little more about making quantitative and qualitative observations. A set 
of observations about a particular phenomenon is called data. Scientist use many techniques to analyze 
and interpret data. Data analysis produces organized data that is more conducive to seeing regularities 
and drawing conclusions. Making tables and graphs of data are two of the most useful techniques in data 
analysis. 



Natural Laws are Statements of Repeated Data Patterns 

Around the year 1800, Jacque Charles and other scientists were working with gases to, among other reasons, 
improve the design of the hot air balloon. These scientists found, after many, many tests, that patterns 
and regularities existed in the observations on gas behavior. If the temperature of the gas increased, the 
volume of the gas increased. This is known as a natural law. A natural law is a relationship that exists 
between variables in a group of data. Natural laws describe the patterns we see in large amounts of data. 
These laws have withstood the test of time because they have been based on repeated observation with no 
known exceptions. 




Figure 3.3: Scuba Divers. 



Around the same time as Charles was working with hot air balloons, another scientist, names J.W. Henry 
was doing experiments trying to find a pattern between the pressure of a gas and the amount of the gas that 
dissolved in water. Henry found that when one of these variables increased, the other variable increased in 
the same proportion. Have you ever gone scuba diving (Figure 3.3)? Scuba Divers learn about a problem 
known as "the Bends" when they are being trained. As scuba divers dive deeper, the increased pressure of 
the breathing air causes more nitrogen to be dissolved in the diver's blood. Coming up too quickly from 
a dive causes the pressure to decrease rapidly and therefore, the nitrogen to leave the blood quickly which 
leads to "the bends." Henry's Law is called a natural law because it indicates a relationship (regularity) 
between gas pressure and the amount of dissolved nitrogen. 



Ill 



www.ckl2.org 



A Hypothesis is a Tentative Explanation 

When scientists develop a description of the nature of matter to explain observations (including natural 
laws), the first attempt at an explanation is often referred to as a hypothesis. A hypothesis is your educated 
or best guess as to the nature of matter that causes those observations. The requirements for a hypothesis 
are only that the hypothesis explains ALL the observations and that it is possible to make an observation 
that will refute the hypothesis. 

The hypothesis must be testable. The test of a hypothesis is called an experiment. If the results of the 
experiment contradict the hypothesis, the hypothesis is rejected and a new hypothesis is formulated. The 
results of the experiment are now included in the observations list and the new hypothesis must explain 
this new observation as well as all the previous observations. If the result of the experiment supports the 
hypothesis, more tests are still required. Hypotheses are not proven by testing . . . they are merely 
supported or contradicted. 

A Theory is an Explanation of a Law 

As stated earlier in this section, a law describes a pattern of data that is observed with no known exception. 
A theory is a possible explanation for a law. In science, theories can either be descriptive (qualitative) or 
mathematical (quantitative), but because they explain the patterns described in the law, theory can be 
used to predict future events. On a popular television show, mathematical theories are used to analyze 
and describe behavior in order to predict future events. Hypotheses that have survived many supportive 
tests are often called theories. Theories have a great deal more supportive testing behind them than do 
hypotheses. 

Let's put it together. The Law of Conservation of Mass that you learned earlier stated that matter cannot 
be created nor destroyed. For example, in the reaction below, you will see that there are 28g + 6g = 34 g 
of reactants (before the arrow) and 34 g of products (after the arrow). 

N 2 ( g ) + 3 H 2 (g) -» 2 NH 3(g) 

28g 6g 34g 

Mass must remain constant from the start of a reaction to completion. This law was the result of many 
quantitative experiments done by John Dalton and others in the early part of the 1800s. Dalton had 
formulated many hypotheses surrounding his vision of how the elements worked together, how compounds 
formed, and how chemical reactions would take place maintaining this mass from beginning to end. Even- 
tually, in 1803, Dalton was able to propose the atomic theory which was an explanation for this law. 

Sample Question: What distinguishes a law from a theory? 

Solution: A law is an observation of nature; a theory is a possible explanation of the law. 

Models Developed to Aid in Understanding 

Scientists often use models when they need a way to communicate their understanding of what might be 
very small (such as an atom or molecule) or very large (such as the universe). A model is another way to 
express a theory. 

If you were asked to determine the contents of a box that cannot be opened, you would do a variety of 
experiments in order to develop an idea (or a model) of what the box contains. You would probably shake 
the box, perhaps put magnets near it and/or determine its mass. When you completed your experiments, 

www.ckl2.org 112 



you would develop an idea of what is inside; that is, you would make a model of what is inside a box that 
cannot be opened. 

A good example of how a model is useful to scientists is the kinetic molecular theory. The theory can be 
defined in statements, but it becomes much more easily understood if representations of the particles in 
their three phases are drawn. 

Another example is how models were used to explain the development of the atomic theory. As you will 
learn in a later chapter, the idea of the concept of an atom changed over many years. In order to understand 
each of the different theories of the atom according to the various scientists, models were drawn, and more 
easily understood. 

Lesson Summary 

• A natural law is an observation, or a description of a large amount of reproducible data. 

• A hypothesis is a early attempt at an explanation for data. 

• A theory is used to explain a law or to explain a series of facts/events. 

• Theories can use qualitative analogies or models to describe results. 

Review Questions 

1. Jack performed an experiment where he measured the masses of two different reactants and the 
resulting product. His results are shown in the equation below. What law is Jack demonstrating in 
his experiment? 

10* 20# 30g 

(a) law of constant composition 

(b) law of combining volumes 

(c) law of conservation of mass 

(d) law of conservation of energy 

(e) law of multiple proportions 

2. Sugar dissolves in water. What kind of a statement is this? 

(a) a hypothesis 

(b) a law 

(c) a theory 

(d) a rule 

(e) all of the above 

3. Draw a model to represent the difference between a solid, a liquid, and a gas. In your model, use 
symbols to represent the molecules that are present in each state. The model should then show how 
the molecules exist in each state. 

Further Reading / Supplemental Links 

• http : //en . wikipedia . org/wiki/Category : Chemistry_theories 

• http : //en . wikipedia . org/wiki/Polytetraf luoroethylene 

113 www.ckl2.org 



Vocabulary 

natural laws A description of the patterns observed in the large amounts of data. 

hypothesis An educated guess as to what is going to happen in the experiment. 

theory Used to explain a law or to explain a series of facts/events. 

law of conservation of mass Matter cannot be created nor destroyed. 

model A description, graphic, or 3-D representation of theory used to help enhance understanding. 

scientific method The method of deriving the theories from hypotheses and laws through experimen- 
tation and observation. 



Image Sources 



(1) http : //www . katrina . noaa . gov/satellite/satellite . html. 

(2) Scuba Divers.. 

(3) Richard Parsons. . CCBYSA. 



www.ckl2.org 114 



Chapter 4 
Atomic Theory 



4.1 Early Development of a Theory 

Lesson Objectives 

• Give a short history of the Concept of the atom. 

• State the Law of Definite Proportions. 

• State the Law of Multiple Proportions. 

• State Dalton's Atomic Theory, and explain its historical development. 

Introduction 

You learned earlier how all matter in the universe is made out of tiny building blocks called atoms. The 
concept of the atom is accepted by all modern scientists, but when atoms were first proposed about 2500 
years ago, ancient philosophers laughed at the idea. It has always been difficult to convince people of 
the existence of things that are too small to see. There are many observations that are made on atoms, 
however, that do not involve actually seeing the atom itself and science is about observing and devising 
a theory to explain why those observations occur. We will spend some time considering the evidence 
(observations) that convince scientists of the existence of atoms. 

Democritus and the "Atom" 

Before we discuss the experiments and evidence which have, over the years, convinced scientists that matter 
is made up of atoms, it's only fair to give credit to the man who proposed "atoms" in the first place. 2,500 
years ago, early Greek philosophers believed the entire universe was a single, huge, entity. In other words, 
"everything was one." They believed that all objects, all matter, and all substance were connected as a 
single, big, unchangeable "thing." Now you're probably disturbed by the word unchangeable. Certainly 
you've seen the world around you change, and those early Greek philosophers must have too. Why, then, 
would they think that the universe was unchangeable? Well, strange as it may sound to you today, back 
then the generally accepted theory was that the world didn't change - it just looked like it did. In other 
words, Greek philosophers believed that all change (and all motion) was an illusion. It was all in your 
head! Compared to this crazy idea, the atom is looking pretty good, isn't it? 

One of the first people to propose "atoms" was a man known as Democritus (Figure 4.1). Democritus 

115 www.ckl2.org 




Figure 4.1: Democritus was known as "The Laughing Philosopher." It's a good thing he liked to laugh, 
because most other philosophers were laughing at his theories. 

didn't like the idea that life was an illusion any more than you probably do. As an alternative, he suggested 
that the world did change, and he explained this change by proposing atomos or atomon - tiny, indivisible, 
solid objects making up all matter in the universe. 

Democritus then reasoned that changes occur when the many atomos in an object were reconnected or 
recombined in different ways. Democritus even extended his theory, suggesting that there were different 
varieties of atomos with different shapes, sizes, and masses. He thought, however, that shape, size and 
mass were the only properties differentiating the different types of atomos. According to Democritus, other 
characteristics, like color and taste, did not reflect properties of the atomos themselves, but rather, resulted 
from the different ways in which the atomos were combined and connected to one another (Figure 4.2). 




Figure 4.2: Democritus believed that properties like color depended on how the atomos were connected to 
each other, and not on the atomos themselves. Interestingly, Democritus was partially right - the green 
emerald and the red ruby both contain atoms of aluminum, oxygen, and chromium. The emerald, however, 
also contains silicon and beryllium atoms. 

Even though the idea of the atomos seems much more reasonable than trying to explain experience as an 
www.ckl2.org 116 



illusion, the early Greek philosophers didn't like it, and they didn't for the following reason. If all matter 
consists of tiny atomos that float around, bang into each other, and connect together in various ways, 
these atomos must be floating in something. But what? Well, according to Democritus, the atoms floated 
around in a void (empty space or "nothingness"). That does seem a bit strange, doesn't it? Certainly, if 
you pound your fist on the desk in front of you, it doesn't feel like there's any empty space in it. What's 
more, Greek philosophers thought that empty space was illogical. In order to exist, they argued, nothing 
must be something, meaning nothing wasn't nothing, but that's a contradiction. Their arguments got 
quite confusing, but the end result was that Greek philosophers dismissed Democritus' theory entirely. 
Sadly, it took over two millennia before the theory of atomos (or "atoms," as they're known today) was 
fully appreciated. 

Greek Philosophers Didn't Experiment 




Figure 4.3: Greek philosophers liked to think - they didn't, however, like to experiment all that much. 

Early Greek philosophers disliked Democritus' theory of atomos because they believed a void, or complete 
"nothingness," was illogical. To ancient thinkers, a theory that went against "logic" was far worse than a 
theory that went against experience or observation. That's because Greek philosophers truly believed that, 
above all else, our understanding of the world should rely on "logic." In fact, they argued that the world 
couldn't be understood using our senses at all, because our senses could deceive us (these were, of course, 
the same people who argued that all change in the world was an illusion). Therefore, instead of relying 
on observation, Greek philosophers tried to understand the world using their minds and, more specifically, 
the power of reason. 

Unfortunately, when Greek philosophers applied reason to Democritus' theory, their arguments were in- 
consistent. Democritus' void had to be "something" to exist, but at the same time it had to be "nothing" 
to be a void. Greek philosophers were not willing to accept the idea that "nothing" could be "something" 
- that seemed illogical. 

Today, we call these contradictions (such as "nothing" is "something") paradoxes. Science is full of para- 
doxes. Sometimes these paradoxes result when our scientific theories are wrong or incomplete, and some- 
times they result because we make bad assumptions about what's "logical" and what isn't. In the case of 

117 www.ckl2.org 



the void, it turns out that "nothing" really can exist, so in a way, "nothing" is "something." 

So how could the Greek philosophers have known that Democritus had a good idea with his theory of 
"atomos?" It would have taken is some careful observation and a few simple experiments. Now you might 
wonder why Greek philosophers didn't perform any experiments to actually test Democritus' theory. The 
problem, of course, was that Greek philosophers didn't believe in experiments at all. Remember, Greek 
philosophers didn't trust their senses, they only trusted the reasoning power of the mind (Figure 4.3). 

Alchemists Experimented But Didn't Seek Explanation 




Figure 4.4: Yet another alchemist searching for the philosopher's stone. Notice how the alchemists used a 
lot of experimental techniques. It's too bad they were only interested in making gold! 

As you learned in the last section, the early Greek philosophers tried to understand the nature of the world 
through reason and logic, but not through experiment and observation. As a result, they had some very 
interesting ideas, but they felt no need to justify their ideas based on life experiences. In a lot of ways, you 
can think of the Greek philosophers as being "all thought and no action." It's truly amazing how much they 
achieved using their minds, but because they never performed any experiments, they missed or rejected a 
lot of discoveries that they could have made otherwise. 

Some of the earliest experimental work was done by the alchemists (Figure 4.4). Remember that the 
alchemists were extremely interested in discovering the "philosopher's stone", which could turn common 
metals into gold. Of course, they also dabbled in medicines and cures, hoping to find "the elixir of life", 
and other such miraculous potions. On the other hand, alchemists were not overly concerned with any 
deep questions about the nature of the world. In contrast to the Greek philosophers, you can think of 



www.ckl2.org 



118 



alchemists as being "all action and no thought." Alchemists experimented freely with everything that they 
could find. In general, though, they didn't think too much about their results and what their results might 
tell them about the world. Instead, they were only interested in whether or not they had made gold. 

To be fair, there were some alchemists who tried to use results from past experiments to help suggest future 
experiments. Nevertheless, alchemy always had very materialistic goals in mind - goals like producing gold 
and living forever. Alchemists were not troubled by philosophical questions like "what is the universe made 
of?" - they didn't really care unless they thought it would somehow help them find the "philosopher's stone" 
or the "elixir of life." 

What you've probably noticed by reading about the Greek philosophers and the alchemists is that the 
history of science is ironic. Greek philosophers asked deep questions about the universe but didn't believe 
in any of the experiments that might have led them to the answers. In contrast, alchemists believed in 
experimentation but weren't interested in what the experiments might tell them in terms of the nature 
of the world. Unbelievably, it took over 2000 years to put the questions asked by the Greek philosophers 
together with the experimental tools developed by the alchemists. The result was significant progress in 
our understanding of nature and the universe, and that's what we'll learn about next. 

Dalton's Atomic Theory 

Let's begin our discussion of Dalton's atomic theory by considering a simple, but important experiment that 
suggested matter might be made up of atoms. In the late 1700's and early 1800's, scientists began noticing 
that when certain substances, like hydrogen and oxygen, were combined to produce a new substance, like 
water, the reactants (hydrogen and oxygen) always reacted in the same proportions by mass. In other 
words, if 1 gram of hydrogen reacted with 8 grams of oxygen, then 2 grams of hydrogen would react 
with 16 grams of oxygen, and 3 grams of hydrogen would react with 24 grams of oxygen. Strangely, the 
observation that hydrogen and oxygen always reacted in the "same proportions by mass" wasn't special. 
In fact, it turned out that the reactants in every chemical reaction reacted in the same proportions by 
mass. Take, for example, nitrogen and hydrogen, which react to produce ammonia (a chemical you've 
probably used to clean your house). 1 gram of hydrogen will react with 4.7 grams of nitrogen, and 2 grams 
of hydrogen will react with 9.4 grams of nitrogen. Can you guess how much nitrogen would react with 
3 grams of hydrogen? 

Scientists studied reaction after reaction, but every time the result was the same. The reactants always 
reacted in the "same proportions by mass" or in what we call "definite proportions." As a result, scientists 
proposed the Law of Definite Proportions (Figure 4.5). This law states that: 

In a given type of chemical substance, the elements are always combined in the same pro- 
portions by mass. 

Earlier, you learned that an "element" is a grouping in which there is only one type of atom - of course, 
when the Law of Definite Proportions was first discovered, scientists didn't know about atoms or elements, 
so the law was stated slightly differently. We'll stick with this modern version, though, since it's easiest to 
understand. 

The Law of Definite Proportions applies when elements are reacted together to form the same product. 
Therefore, while the Law of Definite Proportions can be used to compare two experiments in which hydrogen 
and oxygen react to form water, the Law of Definite Proportions can not be used to compare one experiment 
in which hydrogen and oxygen react to form water, and another experiment in which hydrogen and oxygen 
react to form hydrogen peroxide (peroxide is another material that can be made from hydrogen and oxygen) . 

While scientists around the turn of the 18 century weren't making a lot of peroxide, a man named John 
Dalton was experimenting with several reactions in which the reactant elements formed more than one type 
of product, depending on the experimental conditions he used (Figure 4.6). One common reaction that he 

119 www.ckl2.org 



B 


B 


B 


B 


B 


B 


B 


B 



C 


c 


c 


c 


c 


c 


c 


c 


c 





B 


B 


B 


B 


A 
A 




B 


B 


B 


B 


B 


B 


B 


B 




B 


B 


B 


B 





C 


c 


c 


c 


C 


c 


c 


c 


c 


C 


c 


c 


c 


c 




c 


c 


c 


c 



Figure 4.5: If of reacts with of , then by the Law of Definite Proportions, of must react with of . If of 
reacts with of , then by the Law of Conservation of Mass, they must produce of . Similarly, when of react 
with of , they must produce of . 




Figure 4.6: John Dalton was a thinker, but he was also an experimenter. 



www.ckl2.org 



120 



studied was the reaction between carbon and oxygen. When carbon and oxygen react, they produce two 
different substances - we'ii caff these substances "A" and "B." It turned out that, given the same amount of 
carbon, forming B always required exactly twice as much oxygen as forming A. In other words, if you can 
make A with 3 grams of carbon and 4 grams of oxygen, B can be made with the same 3 grams of carbon, 
but with 8 grams oxygen. Dalton asked himself - why does B require 2 times as much oxygen as A? Why 
not 1.21 times as much oxygen, or 0.95 times as much oxygen? Why a whole number like 2? 

The situation became even stranger when Dalton tried similar experiments with different substances. For 
example, when he reacted nitrogen and oxygen, Dalton discovered that he could make three different 
substances - we'll call them "C," "D," and "£." As it turned out, for the same amount of nitrogen, D 
always required twice as much oxygen as C. Similarly, E always required exactly four times as much 
oxygen as C. Once again, Dalton noticed that small whole numbers (2 and 4) seemed to be the rule. 

Dalton used his experimental results to propose The Law of Multiple Proportions: 

When two elements react to form more than one substance, the different masses of one 
element (like oxygen) that are combined with the same mass of the other element (like 
nitrogen) are in a ratio of small whole numbers. 

Dalton thought about his Law of Multiple Proportions and tried to find some theory that would explain it. 
Dalton also knew about the Law of Definite Proportions and the Law of Conservation of Mass (Remember 
that the Law of Conservation of Mass states that mass is neither created nor destroyed), so what he really 
wanted was a theory that would explain all three of these laws using a simple, plausible model. One way to 
explain the relationships that Dalton and others had observed was to suggest that materials like nitrogen, 
carbon and oxygen were composed of small, indivisible quantities which Dalton called "atoms" (in reference 
to Democritus' original idea). Dalton used this idea to generate what is now known as Dalton' s Atomic 
Theory* 

Dalton's Atomic Theory 

1. Matter is made of tiny particles called atoms. 

2. Atoms are indivisible. During a chemical reaction, atoms are rearranged, but they do not break 
apart, nor are they created or destroyed. 

3. All atoms of a given element are identical in mass and other properties. 

4. The atoms of different elements differ in mass and other properties. 

5. Atoms of one element can combine with atoms of another element to form "compounds" - new, 
complex particles. In a given compound, however, the different types of atoms are always present in 
the same relative numbers. 

*Some people think that Dalton developed his Atomic Theory before stating the Law of Multiple Propor- 
tions, while others argue that the Law of Multiple Proportions, though not formally stated, was actually 
discovered first. In reality, Dalton was probably contemplating both concepts at the same time, although 
it is hard to tell from his laboratory notes. 

Lesson Summary 

• 2,500 years ago, Democritus suggested that all matter in the universe was made up of tiny, indivisible, 
solid objects he called "atomos." 

• Democritus believed that there were different types of "atomos" which differed in shape, size, and 
mass. 

• Other Greek philosophers disliked Democritus' "atomos" theory because they felt it was illogical. 
Since they didn't believe in experiments, though, they had no way to test the "atomos" theory. 

121 www.ckl2.org 



• Alchemists experimented and developed experimental techniques, but they were more interested in 
making gold than they were in understanding the nature of matter and the universe. 

• The Law of Definite Proportions states that in a given chemical substance, the elements are always 
combined in the same proportions by mass. 

• The Law of Multiple Proportions states that when two elements react to form more than one sub- 
stance, the different masses of one element that are combined with the same mass of the other element 
are in a ratio of small whole numbers. 

• Dalton used the Law of Definite Proportions, the Law of Multiple Proportions, and The Law of 
Conservation of Mass to propose his Atomic Theory. 

• Dalton's Atomic Theory states: 

1. Matter is made of tiny particles called atoms. 

2. Atoms are indivisible. During a chemical reaction, atoms are rearranged, but they do not break 
apart, nor are they created or destroyed. 

3. All atoms of a given element are identical in mass and other properties. 

4. The atoms of different elements differ in mass and other properties. 

5. Atoms of one element can combine with atoms of another element to form "compounds" - new 
complex particles. In a given compound, however, the different types of atoms are always present in 
the same relative numbers. 

Review Questions 

1. It turns out that a few of the ideas in Dalton's Atomic Theory aren't entirely correct. Are inaccurate 
theories an indication that science is a waste of time? 

2. Suppose you are trying to decide whether to wear a sweater or a T-shirt. To make your decision, you 
phone two friends. The first friend says, "Wear a sweater, because I've already been outside today, 
and it's cold." The second friend, however, says, "Wear a T-shirt. It isn't logical to wear a sweater in 
July." Would you decide to go with your first friend, and wear a sweater, or with your second friend, 
and wear a T-shirt? Why? 

3. Decide whether each of the following statements is true or false. 

(a) Democritus believed that all matter was made of "atomos." 

(b) Democritus also believed that there was only one kind of "atomos." 

(c) Most early Greek scholars thought that the world was "ever-changing." 

(d) If the early Greek philosophers hadn't been so interested in making gold, they probably would 
have liked the idea of the "atomos." 

4. Match the person, or group of people, with their role in the development of chemistry. 

(a) Early Greek philosophers - a. suggested that all matter was made up of tiny, indivisible objects 

(b) alchemists - b. tried to apply logic to the world around them 

(c) John Dalton - c. suggested that all matter was made up of tiny, indivisible objects 

(d) Democritus - d. were primarily concerned with finding ways to turn common metals into gold 

5. Early Greek philosophers felt that Democritus "atomos" theory was illogical because: 

(a) no matter how hard they tried, they could never break matter into smaller pieces. 

(b) it didn't help them to make gold. 

(c) sulfur is yellow and carbon is black, so clearly "atomos" must be colored. 

(d) empty space is illogical because it implies that nothing is actually something. 

6. Which Law explains the following observation: carbon monoxide can be formed by reacting 12 grams 
of carbon with 16 grams of oxygen? To form carbon dioxide, however, 12 grams of carbon must react 
with 32 grams of oxygen. 

www.ckl2.org 122 



7. Which Law explains the following observation: carbon monoxide can be formed by reacting 12 grams 
of carbon with 16 grams of oxygen? It can also be formed by reacting 24 grams of carbon with 
32 grams of oxygen. 

8. Which Law explains the following observation: 28 grams of carbon monoxide are formed when 
12 grams of carbon reacts with 16 grams of oxygen? 

9. Which Law explains the following observations: when 12 grams of carbon react with 4 grams of 
hydrogen, they produce methane, and there is no carbon or hydrogen left over at the end of the 
reaction? If, however, 11 grams of carbon react with 4 grams of hydrogen, there is hydrogen left over 
at the end of the reaction. 

10. Which of the following is not part of Dalton's Atomic Theory? 

(a) matter is made of tiny particles called atoms. 

(b) during a chemical reaction, atoms are rearranged. 

(c) during a nuclear reaction, atoms are split apart. 

(d) all atoms of a specific element are the same. 

Calculations 

11. Consider the following data: 3.6 grams of boron react with 1.0 grams of hydrogen to give 4.6 grams 
of BH3. How many grams of boron would react with 2.0 grams of hydrogen? 

12. Consider the following data: 12 grams of carbon and 4 grams of hydrogen react to give 16 grams of 
"compound A." 24 grams of carbon and 6 grams of hydrogen react to give 30 grams of "compound 
5." Are compound A and compound B the same? Why or why not? 

Vocabulary 

atomos (atomon) Democritus' word for the tiny, indivisible, solid objects that he believed made up all 

matter in the universe. 
void Another word for empty space. 

paradox Two statements that seem to be true, but contradict each other. 

law of definite proportions In a given chemical substance, the elements are always combined in the 
same proportions by mass. 

law of multiple proportions When two elements react to form more than one substance, the different 
masses of one element that are combined with the same mass of the other element are in a ratio of 
small whole numbers. 

4.2 Further Understanding of the Atom 

Lesson Objectives 

• Explain the experiment that led to Thomson's discovery of the electron. 

• Describe Thomson's "plum pudding" model of the atom. 

• Describe Rutherford's Gold Foil experiment, and explain how this experiment proved that the "plum 
pudding" model of the atom was incorrect. 

123 www.ckl2.org 



Introduction 

In the last lesson, you learned about the atom, and the early experiments that led to the development of 
Dalton's Atomic Theory. But Dalton's Atomic Theory isn't the end of the story. Do you remember the 
scientific method introduced in early on in the book? Chemists using the scientific method make careful 
observations and measurements and then use these measurements to propose theories. That's exactly what 
Dalton did. Dalton used the following observations: 

1. Mass is neither created nor destroyed during a chemical reaction. 

2. Elements always combine in the same proportions by mass when they form a given compound. 

3. When elements form more than one compound, the different masses of one element that are combined 
with the same mass of the other element are in a ratio of small whole numbers. 

With these observations (which came from careful measurement), Dalton proposed his Atomic Theory - a 
model which suggested how the underlying structure of matter might lead to the three observations above. 
The scientific method, though, doesn't stop once a theory has been proposed. Instead, the theory should 
suggest new experiments that can be performed to test whether or not the original theory is accurate and 
complete. 

Dalton's Atomic Theory held up well to a lot of the different chemical experiments that scientists performed 
to test it. In fact, for almost 100 years, it seemed as if Dalton's Atomic Theory was the whole truth. 
However, in 1897, a scientist named J. J. Thompson conducted some research which suggested that Atomic 
Theory wasn't the entire story. As it turns out, Dalton had a lot right. He was right in saying matter is 
made up of atoms; he was right in saying there are different kinds of atoms with different mass and other 
properties; he was "almost" right in saying atoms of a given element are identical; he was right in saying 
during a chemical reaction, atoms are merely rearranged; he was right in saying a given compound always 
has atoms present in the same relative numbers. But he was WRONG in saying atoms were indivisible 
or indestructible. As it turns out, atoms are divisible. In fact, atoms are composed of smaller subatomic 
particles. We'll talk about the discoveries of these subatomic particles next. 

Thomson Discovered Electrons Were Part of the Atom 

In the mid- 1800s, scientists were beginning to realize that the study of chemistry and the study of electricity 
were actually related. First, a man named Michael Faraday showed how passing electricity through mixtures 
of different chemicals could cause chemical reactions. Shortly after that, scientists found that by forcing 
electricity through a tube filled with gas, the electricity made the gas glow! Scientists didn't, however, 
understand the relationship between chemicals and electricity until a British physicist named J.J. Thomson 
(Figure 4.7) began experimenting with what is known as a cathode ray tube. 

Figure ?? shows a basic diagram of a cathode ray tube like the one J. J. Thomson would have used. A 
cathode ray tube is a small glass tube with a cathode (a negatively charged metal plate) and an anode (a 
positively charged metal plate) at opposite ends. By separating the cathode and anode by a short distance, 
the cathode ray tube can generate what are known as "cathode rays" - rays of electricity that flowed from 
the cathode to the anode, J. J. Thomson wanted to know what cathode rays were, where cathode rays 
came from and whether cathode rays had any mass or charge. The techniques that J. J. Thomson used 
to answer these questions were very clever and earned him a Nobel Prize in physics. First, by cutting a 
small hole in the anode J. J. Thomson found that he could get some of the cathode rays to flow through 
the hole in the anode and into the other end of the glass cathode ray tube. Next, J. J. Thomson figured 
out that if he painted a substance known as "phosphor" onto the far end of the cathode ray tube, he could 
see exactly where the cathode rays hit because the cathode rays made the phosphor glow. 

www.ckl2.org 124 




Figure 4.7: A portrait of J. J. Thomson. 



A cathode-ray tube. (Source: Sharon Bewick. CC-BY-SA) 



electricity passes from the 
cathode to the anode 



a small bit of electricity 
passes through the hole 
in the anode 



cathode 




phophor coating 



J. J. Thomson must have suspected that cathode rays were charged, because his next step was to place a 
positively charged metal plate on one side of the cathode ray tube and a negatively charged metal plate 
on the other side of the cathode ray tube, as shown in (Figure ??). The metal plates didn't actually 
touch the cathode ray tube, but they were close enough that a remarkable thing happened! The flow of 
the cathode rays passing through the hole in the anode was bent upwards towards the positive metal plate 
and away from the negative metal plate. 

A cathode ray was attracted to the positively charged metal plate placed above the tube, and repelled 
from negatively charged metal plate placed below the tube. (Source: Sharon Bewick. CC-BY-SA) 



125 



www.ckl2.org 



positively charged 

metal plate 

+ 



^^ 



Hi 



negatively charged 
metal plate 




In other words, instead of the phosphor glowing directly across from the hole in the anode (as in Figure 
??), the phosphor now glowed at a spot quite a bit higher in the tube (as in Figure ??). 

J. J. Thomson thought about his results for a long time. It was almost as if the cathode rays were attracted 
to the positively charged metal plate above the cathode ray tube, and repelled from the negatively charged 
metal plate below the cathode ray tube. J. J. Thomson knew that charged objects are attracted to and 
repelled from other charged objects according to the rule: opposites attract, likes repel. This means that 
a positive charge is attracted to a negative charge, but repelled from another positive charge. Similarly, 
a negative charge is attracted to a positive charge, but repelled from another negative charge. Using the 
"opposites attract, likes repel" rule, J. J. Thomson argued that if the cathode rays were attracted to the 
positively charged metal plate and repelled from the negatively charged metal plate, they themselves must 
have a negative charge! 

J. J. Thomson then did some rather complex experiments with magnets, and used his results to prove that 
cathode rays were not only negatively charged, but also had mass. Remember that anything with mass is 
part of what we call matter. In other words, these cathode rays must be the result of negatively charged 
"matter" flowing from the cathode to the anode. But there was a problem. According to J. J. Thomson's 
measurements, either these cathode rays had a ridiculously high charge, or else had very, very little mass - 
much less mass than the smallest known atom. How was this possible? How could the matter making up 
cathode rays be smaller than an atom if atoms were indivisible? J. J. Thomson made a radical proposal: 
maybe atoms are divisible. J. J. Thomson suggested that the small, negatively charged particles making 
up the cathode ray were actually pieces of atoms. He called these pieces "corpuscles," although today we 
know them as "electrons." Thanks to his clever experiments and careful reasoning J. J. Thomson is credited 
with the discovery of the electron. 



www.ckl2.org 



126 



Protons Were Thought to Exist but Discovered Much Later 

In the last section, we learned that atoms are, in fact, divisible, and that one of the subatomic particles 
making up an atom is a small, negatively charged entity called an "electron." Now imagine what would 
happen if atoms were made entirely of electrons. First of all, electrons are very, very small; in fact, electrons 
are about 2000 times smaller than the smallest known atom, so every atom would have to contain a whole 
lot of electrons. But there's another, even bigger problem: electrons are negatively charged. Therefore, if 
atoms were made entirely out of electrons, atoms would be negatively charged themselves... and that would 
mean all matter was negatively charged as well. 

Of course, matter isn't negatively charged. In fact, most matter is what we call neutral - it has no charge 
at all. If matter is composed of atoms, and atoms are composed of negative electrons, how can matter 
be neutral? The only possible explanation is that atoms consist of more than just electrons. Atoms must 
also contain some type of positively charged material which balances the negative charge on the electrons. 
Negative and positive charges of equal size cancel each other out, just like negative and positive numbers 
of equal size. What do you get if you add +1 and -1? You get 0, or nothing. That's true of numbers, 
and that's also true of charges. If an atom contains an electron with a -1 charge, but also some form of 
material with a +1 charge, overall the atom must have a (+1) + (-1) = charge - in other words, the 
atom must be neutral, or have no charge at all. 

Based on the fact that atoms are neutral, and based on J. J. Thomson's discovery that atoms contain 
negative subatomic particles called "electrons," scientists assumed that atoms must also contain a positive 
substance. It turned out that this positive substance was another kind of subatomic particle, known 
as the "proton." Although scientists knew that atoms had to contain positive material, protons weren't 
actually discovered, or understood, until quite a bit later. 

Thomson's Model of the Atom 



Figure 4.8: A plum pudding and Thomson's "plum-pudding" model for the atom. Notice how the "plums" 
are the negatively charged electrons, while the positive charge is spread throughout the entire pudding 
batter. 

When Thomson discovered the negative electron, he realized that atoms had to contain positive material 
as well - otherwise they wouldn't be neutral overall. As a result, Thomson formulated what's known as 
the "plum pudding" model for the atom. According to the "plum pudding" model, the negative electrons 
were like pieces of fruit and the positive material was like the batter or the pudding. This made a lot of 
sense given Thomson's experiments and observations. Thomson had been able to isolate electrons using 

127 www.ckl2.org 



a cathode ray tube; however he had never managed to isolate positive particles. As a result, Thomson 
theorized that the positive material in the atom must form something like the "batter" in a plum pudding, 
while the negative electrons must be scattered through this "batter." (If you've never seen or tasted a plum 
pudding, you can think of a chocolate chip cookie instead. In that case, the positive material in the atom 
would be the "batter" in the chocolate chip cookie, while the negative electrons would be scattered through 
the batter like chocolate chips.) 

Figure 4.8 shows a "plum pudding" and a "plum pudding" model for the atom. Notice how easy it would 
be to pick the pieces of fruit out of a plum pudding. On the other hand, it would be a lot harder to 
pick the batter out of the plum pudding, because the batter is everywhere. If an atom were similar to a 
plum pudding in which the electrons are scattered throughout the "batter" of positive material, then you'd 
expect it would be easy to pick out the electrons, but a lot harder to pick out the positive material. 

Everything about Thomson's experiments suggested the "plum pudding" model was correct - but accord- 
ing to the scientific method, any new theory or model should be tested by further experimentation and 
observation. In the case of the "plum pudding" model, it would take a man named Ernest Rutherford to 
prove it wrong. Rutherford and his experiments will be the topic of the next section. 



Rutherford's Model of the Atom 




Disproving Thomson's "plum pudding" model began with the discovery that an element known as uranium 
emitted positively charged particles called alpha particles as it underwent radioactive decay. Radioactive 
decay occurs when one element decomposes into another element. It only happens with a few very unstable 
elements. This involves some difficult concepts so, for now, just accept the fact that uranium decays and 
emits alpha particles in the process. Alpha particles themselves didn't prove anything about the structure 
of the atom. In fact, a man named Ernest Rutherford proved that alpha particles were nothing more than 
helium atoms that had lost their electrons (Figure ??). Think about why an atom that has lost electrons 
will have a positive charge. Alpha particles could, however, be used to conduct some very interesting 
experiments. 

Ernest Rutherford was fascinated by all aspects of alpha particles. For the most part, though, he seemed 
to view alpha particles as tiny bullets that he could use to fire at all kinds of different materials. One 
experiment in particular, however, surprised Rutherford, and everyone else. Rutherford found that when he 

www.ckl2.org 128 



fired alpha particles at a very thin piece of gold foil, an interesting thing happened (Figure ??). Almost 
all of the alpha particles went straight through the foil as if they'd hit nothing at all. Every so often, 
though, one of the alpha particles would be deflected slightly as if it had bounced off of something hard. 
Even less often, Rutherford observed alpha particles bouncing straight back at the "gun" from which they 
had been fired! It was as if these alpha particles had hit a wall "head-on" and had ricocheted right back 
in the direction that they had come from. 

Rutherford thought that these experimental results were rather odd. Rutherford described firing alpha 
particles at gold foil like shooting a high-powered rifle at tissue paper. Would you ever expect the bullets 
to hit the tissue paper and bounce back at you? Of course not! The bullets would break through the tissue 
paper and keep on going, almost as if they'd hit nothing at all. That's what Rutherford had expected 
would happen when he fired alpha particles at the gold foil. Therefore, the fact that most alpha particles 
passed through didn't shock him. On the other hand, how could he explain the alpha particles that got 
deflected? Even worse, how could he explain the alpha particles that bounced right back as if they'd hit a 
wall? 

Rutherford decided that the only way to explain his results was to assume that the positive matter forming 
the gold atoms was not, in fact, distributed like the batter in plum pudding, but rather, was concentrated 
in one spot, forming a small positively charged particle somewhere in the center of the gold atom. We 
now call this clump of positively charged mass the nucleus. According to Rutherford, the presence of a 
nucleus explained his experiments, because it implied that most alpha particles passed through the gold 
foil without hitting anything at all. Once in a while, though, the alpha particles would actually collide with 
a gold nucleus, causing the alpha particles to be deflected, or even to bounce right back in the direction 
they came from. 



a particles 




nucleus go 



deflected 
Ct, particles 



non-deflected 
a particles 



gold atom 



A short history of the changes in our model of the atom, an image of the plum pudding model, and an 
animation of Rutherford's experiment can be viewed at Plum Pudding and Rutherford Page (http : //www . 
newcastle-schools . org . uk/nsn/chemistry/Rad.ioactivity/Plub /„20Pudding /o20and. / 20Rutherf ord°/ 20Page . 
htm). 



129 



www.ckl2.org 



Rutherford Suggested Electrons "Orbited" 

While Rutherford's discovery of the positively charged atomic nucleus offered insight into the structure of 
the atom, it also led to some questions. According to the "plum pudding" model, electrons were like plums 
embedded in the positive "batter" of the atom. Rutherford's model, though, suggested that the positive 
charge wasn't distributed like batter, but rather, was concentrated into a tiny particle at the center of the 
atom, while most of the rest of the atom was empty space. What did that mean for the electrons? If 
they weren't embedded in the positive material, exactly what were they doing? And how were they held in 
the atom? Rutherford suggested that the electrons might be circling or "orbiting" the positively charged 
nucleus as some type of negatively charged cloud, but at the time, there wasn't much evidence to suggest 
exactly how the electrons were held in the atom. 

Despite the problems and questions associated with Rutherford's experiments, his work with alpha particles 
definitely seemed to point to the existence of an atomic "nucleus." Between J. J. Thomson, who discovered 
the electron, and Rutherford, who suggested that the positive charges in an atom were concentrated at the 
atom's center, the 1890s and early 1900s saw huge steps in understanding the atom at the "subatomic" (or 
smaller than the size of an atom) level. Although there was still some uncertainty with respect to exactly 
how subatomic particles were organized in the atom, it was becoming more and more obvious that atoms 
were indeed divisible. Moreover, it was clear that the pieces an atom could be separated into negatively 
charged electrons and a nucleus containing positive charges. In the next lesson, we'll look more carefully 
at the structure of the nucleus, and we'll learn that while the atom is made up of positive and negative 
particles, it also contains neutral particles that neither Thomson, nor Rutherford, were able to detect with 
their experiments. 

Lesson Summary 

• Dalton's Atomic Theory wasn't entirely correct. It turns out that atoms can be divided into smaller 
subatomic particles. 

• A cathode ray tube is a small glass tube with a cathode and an anode at one end. Cathode rays flow 
from the cathode to the anode. 

• When cathode rays hit a material known as "phosphor" they cause the phosphor to glow. J. J. 
Thomson used this phenomenon to reveal the path taken by a cathode ray in a cathode ray tube. 

• J. J. Thomson found that the path taken by the cathode ray could be bent towards a positive metal 
plate, and away from a negative metal plate. As a result, he reasoned that the particles in the cathode 
ray were negative. 

• Further experiments with magnets proved that the particles in the cathode ray also had mass. Thom- 
son's measurements indicated, however, that the particles were much smaller than atoms. 

• J. J. Thomson suggested that these small, negatively charged particles were actually subatomic 
particles. We now call them "electrons." 

• Since atoms are neutral, atoms that contain negatively charged electrons must also contain positively 
charged material. 

• According to Thomson's "plum pudding" model, the negatively charged electrons in an atom are like 
the pieces of fruit in a plum pudding, while the positively charged material is like the batter. 

www.ckl2.org 130 



• When Ernest Rutherford fired alpha particles at a thin gold foil, most alpha particles went straight 
through; however, a few were scattered at different angles, and some even bounced straight back. 

• In order to explain the results of his Gold Foil experiment, Rutherford suggested that the positive 
matter in the gold atoms was concentrated at the center of the gold atom in what we now call the 
nucleus of the atom. 

• Rutherford's model of the atom didn't explain where electrons were located in an atom. 

Review Questions 

Concepts 

1. Decide whether each of the following statements is true or false. 

(a) Cathode rays are positively charged. 

(b) Cathode rays are rays of light, and thus they have no mass. 

(c) Cathode rays can be repelled by a negatively charged metal plate. 

(d) J.J. Thomson is credited with the discovery of the electron. 

(e) Phosphor is a material that glows when struck by cathode rays. 

2. Match each observation with the correct conclusion. 

(a) Cathode rays are attracted to a positively charged metal plate. - i. Cathode rays are positively 
charged. - ii. Cathode rays are negatively charged. - iii. Cathode rays have no charge. 

(b) Electrons have a negative charge. - i. atoms must be negatively charged. - ii. atoms must be 
positively charged. - iii. atoms must also contain positive subatomic material. 

(c) Alpha particles fired at a thin gold foil are occasionally scattered back in the direction that 
they came from - i. the positive material in an atom is spread throughout like the "batter" in 
pudding - ii. atoms contain neutrons - iii. the positive charge in an atom is concentrated in a 
small area at the center of the atom. 

3. Alpha particles are: 

(a) Helium atoms that have extra electrons. 

(b) Hydrogen atoms that have extra electrons. 

(c) Hydrogen atoms that have no electrons. 

(d) Electrons. 

(e) Helium atoms that have lost their electrons. 

(f) Neutral helium atoms. 

4. What is the name given to the tiny clump of positive material at the center of an atom? 

5. Choose the correct statement. 

(a) Ernest Rutherford discovered the atomic nucleus by performing experiments with aluminum 
foil. 

(b) Ernest Rutherford discovered the atomic nucleus using a cathode ray tube. 

(c) When alpha particles are fired at a thin gold foil, they never go through. 

(d) Ernest Rutherford proved that the "plum pudding model" was incorrect. 

(e) Ernest Rutherford experimented by firing cathode rays at gold foil. 

6. Answer the following questions: 

(a) Will the charges +2 and -2 cancel each other out? 

(b) Will the charges +2 and -1 cancel each other out? 

131 www.ckl2.org 



(c) Will the charges +1 and +1 cancel each other out? 

(d) Will the charges -1 and +3 cancel each other out? 

(e) Will the charges +9 and -9 cancel each other out? 

7. Electrons are negatively charged metals plates and positively charged metal 

plates? 

8. What was J. J. Thomson's name for electrons? 

9. A "sodium cation" is a sodium atom that has lost one of its electrons. Would the charge on a sodium 
cation be positive, negative or neutral? Would sodium cations be attracted to a negative metal plate, 
or a positive metal plate? Would electrons be attracted to or repelled from sodium cations? 

10. Suppose you have a cathode ray tube coated with phosphor so that you can see where on the tube the 
cathode ray hits by looking for the glowing spot. What will happen to the position of this glowing 
spot if: 

(a) a negatively charged metal plate is placed above the cathode ray tube 

(b) a negatively charged metal plate is placed to the right of the cathode ray tube 

(c) a positively charged metal plate is placed to the right of the cathode ray tube 

(d) a negatively charged metal plate is placed above the cathode ray tube, and a positively charged 
metal plate is placed to the left of the cathode ray tube 

(e) a positively charged metal plate is placed below the cathode ray tube, and a positively charged 
metal plate is also placed to the left of the cathode ray tube. 

Vocabulary 

subatomic particles Particles that are smaller than the atom. The three main subatomic particles are 
electrons, protons and neutrons. 

cathode rays rays of electricity that flow from the cathode to the anode. J.J. Thomson proved that 
these rays were actually negatively charged subatomic particles (or electrons). 

cathode A negatively charged metal plate. 

anode A positively charged metal plate. 

cathode ray tube A glass tube with a cathode and anode, separated by some distance, at one end. 
Cathode ray tubes generate cathode rays. 

phosphor A chemical that glows when it is hit by a cathode ray. 

plum pudding model A model of the atom which suggested that the negative electrons were like plums 
scattered through the positive material (which formed the batter). 

alpha (a) particles Helium atoms that have lost their electrons. They are produced by uranium as it 
decays. 

nucleus The small central core of the atom where most of the mass of the atom (and all of the atoms 
positive charge) is located. 

www.ckl2.org 132 



4.3 Atomic Terminology 

Lesson Objectives 

• Describe the properties of electrons, protons, and neutrons. 

• Define and use an atom's atomic number (Z) and mass number (A). 

• Define an isotope, and explain how isotopes affect an atom's mass, and an element's atomic mass. 

Introduction 

Dalton's Atomic Theory explained a lot about matter, chemicals, and chemical reactions. Nevertheless, it 
wasn't entirely accurate, because contrary to what Dalton believed, atoms can, in fact, be broken apart 
into smaller subunits or subatomic particles. One type of subatomic particle found in an atom is the 
negatively charged electron. Since atoms are neutral, though, they also have to contain positive material. 
At first, scientists weren't sure exactly what this positive material was, or how it existed in the atom. 
Thomson thought it was distributed throughout the atom like batter in a plum pudding. Rutherford, 
however, showed that this was not the case. In his Gold Foil experiment, Rutherford proved that the 
positive substance in an atom was concentrated in a small area at the center of the atom, leaving most the 
rest of the atom as empty space (possibly with a few electrons, or an "electron cloud"). 

Both Thomson's experiments and Rutherford's experiments answered a lot of questions, but they also 
raised a lot of questions, and scientists wanted to know more. How were the electrons connected to the 
rest of the atom? What was the positive material at the center of the atom like? Was it one giant clump 
of positive mass, or could it be divided into smaller parts as well? In this lesson, we'll look at the atom a 
little more closely. 

Electrons, Protons, and Neutrons 

The atom is composed of three different kinds of subatomic particles. First, there are the electrons, which 
we've already talked about, and which J. J. Thomson discovered. Electrons have a negative charge. As 
a result they are attracted to positive objects, and repelled from negative objects, which means that they 
actually repel each other (Figure 4.9). 





Neutral objects neither attract or repel. 



Objects with the same charge repel. 



Figure 4.9: Electrons repel each other because they are both negatively charged. 

Still, most atoms have more than one electron. That's because atoms are big enough to hold many electrons 
without those electrons ever colliding with each other. As you might expect, the bigger the atom, the more 
electrons it contains. 



133 



www.ckl2.org 



Protons are another type of subatomic particle found in atoms. Protons have a positive charge. As a result 
they are attracted to negative objects, and repelled from positive objects. Again, this means that protons 
repel each other (Figure 4.10). Unlike electrons, however, which manage to stake out a 'territory' and 
'defend' it from other electrons, protons are bound together by what are termed strong nuclear forces. 
Therefore, even though they repel each other, protons are forced to group together into one big clump. 
This clump of protons helps to form the nucleus of the atom. Remember, the nucleus of the atom is the 
mass of positive charge at the atom's center. 





Neutral objects neither attract or repel. 



Objects with the same charge repel. 



Figure 4.10: Protons repel each other because they are both positively charged. Despite this repulsion, 
protons are bound together in the atomic nucleus as a result of the strong nuclear force. 

Electrons were the first subatomic particles discovered and protons were the second. There's a third kind of 
subatomic particle, though, known as a neutron, which wasn't discovered until much later. As you might 
have already guessed from its name, the neutron is neutral. In other words, it has no charge whatsoever, 
and is therefore neither attracted to nor repelled from other objects. That's part of the reason why the 
neutron wasn't discovered until long after people knew about electrons and protons - because it has no 
charge, it's really hard to detect. Neutrons are in every atom (with one exception), and they're bound 
together with other neutrons and protons in the atomic nucleus. Again, the binding forces that help to 
keep neutrons fastened into the nucleus are known as strong nuclear forces. 

Before we move on, we must discuss how the different types of subatomic particles interact with each other. 
When it comes to neutrons, the answer is obvious. Since neutrons are neither attracted to, nor repelled 
from objects, they don't really interact with protons or electrons (beyond being bound into the nucleus 
with the protons). Protons and electrons, however, do interact. Using what you know about protons and 
electrons, what do you think will happen when an electron approaches a proton - will the two subatomic 
particles be attracted to each other, or repelled from each other? Here's a hint: "opposites attract, likes 
repel." Electrons and protons have opposite charges (one negative, the other positive), so you'd expect them 
to be attracted to each other and that's exactly what happens (Figure 4.11). 

Viewers journey inside the atom to appreciate its architectural beauty and grasp how atomic structure 
determines chemical behavior. Video on Demand - The World of Chemistry - The Atom (http : //www . 
learner . org/vod/vod_window . html?pid=798) 

Relative Mass and Charge 

Even though electrons, protons, and neutrons are all types of subatomic particles, they are not all the same 
size. When you compare the masses of electrons, protons and neutrons, what you find is that electrons 
have an extremely small mass, compared to either protons or neutrons. On the other hand, the masses of 
protons and neutrons are fairly similar, although technically, the mass of a neutron is slightly larger than 
the mass of a proton. Because protons and neutrons are so much more massive than electrons, almost all 
of the atomic mass in any atom comes from the nucleus, which contains all of the neutrons and protons. 



www.ckl2.org 



134 





Neutral objects neither attract or repel. 



Oppositely charged object? attract. 



Figure 4.11: Protons and electrons are attracted to each other because they have opposite charges. Protons 
are positively charged, while electrons are negatively charged. 

Table 4.1: Masses of the Different Subatomic Particles 



Mass in Grams (g) 



Mass in Atomic Mass Units 
(amu) 



Electron 

Proton 

Neutron 



9.109383 x 1(T 28 
1.6726217 xl0~ 24 
1.6749273 x 10~ 24 



5.485799095 x 10~ 4 

1.0072764669 

1.0086649156 



Table 4.1 gives the masses of electrons, protons, and neutrons. The second column shows the masses of the 
three subatomic particles in grams (which is related to the SI unit kilograms according to the relationship 
1 kg = 1000 g). The third column, however, shows the masses of the three subatomic particles in "atomic 
mass units". Atomic mass units (amu) are useful, because, as you can see, the mass of a proton and the 
mass of a neutron are almost exactly 1.0 in this unit system. We'll discuss atomic mass units in a later 
section. 

Unfortunately, the numbers in Table 4.1 probably don't give you a very good sense of just how big protons 
and neutrons are compared to electrons, so here's a comparison that might help. If an electron were the 
size of a penny, then a proton (or a neutron) would be about the size of a large bowling ball (Figure 4.12). 




Figure 4.12: Electrons are much smaller than protons or neutrons. How much smaller? If an electron was 
the size of a penny, a proton or a neutron would have the mass of a large bowling ball! 

Obviously, if you were told to lift a box containing ^several bowling balls and several pennies,^8vFwou'rcm r 'f 
really care about the pennies, because they wouldn't change the weight of the box all that much. What 
you'd want to know, though, would be the number of bowling balls in the box. That's exactly what 
harmens when scientists trv to fipmre nut the masses of atoms. Thev Hnn't reallv care hnw mamv electrons 



"elementary charge units"* or "elementary charges." Elementary charge units (e) are appealing, because 
the charge on a proton and the charge on an electron are exactly 1.0 in this unit system. 

Table 4.2: Charges on the Different Subatomic Particles 

Charge in Coulombs (C) Mass in Elementary Charges (e) 

Electron -1.6021765 x 10~ 19 -1 

Proton 1.6021765 x 10~ 19 1 

Neutron 



Notice that whether you use Coulombs or elementary charge units, when you ignore the positive and 
negative signs, the charge on the proton and the charge on the electron have the same magnitude. 

Previously, you learned that negative and positive charges of equal magnitude cancel each other out. This 
means that the negative charge on an electron perfectly balances the positive charge on the proton. In 
other words, a neutral atom must have exactly one electron for every proton. If a neutral atom has 1 
proton, it must have 1 electron. If a neutral atom has 2 protons, it must have 2 electrons. If a neutral 
atom has 10 protons, it must have 10 electrons. You get the idea. In order to be neutral, an atom must 
have the same number of electrons and protons, but what kinds of numbers are we talking about? That's 
what we'll look at in the next section. 

• Most scientists don't refer to "elementary charges" as a unit. Nevertheless, if you treat elementary 
charges just like you'd treat any another non-SI unit, like a pound (lb) or a foot (ft), they become a 
lot easier to understand. 

Atomic Number (Z) Identifies the Element 

How do you tell the difference between a bike and a car? What about the difference between a car and a 
unicycle? Take a look at Figure 4.13. 

If you had to make a rule to distinguish between a unicycle, a bike, a car, what would it be? You can't use 
color, because different cars can be different colors and, even worse, a car can be the same color as a bike 
or unicycle. The same goes for weight. While most cars would weigh more than most bikes, which would 
weigh more than most unicycles, that isn't always the case. In fact, that the little grey "Smart Car" in 
Figure 4.13 probably weighs less than a large motorbike. 

What you really need to distinguish between a car, a bike and a unicycle is a property that is the same 
within each category, but different between the categories. A good choice would be the number of wheels. 
All unicycles have one wheel, all bikes have two wheels, and all cars have four wheels. If you count wheels, 
you will most likely never confuse a unicycle with a bike, or a bike with a car (even a motorbike with a 
Smart Car!). In other words, if you know the number of wheels, you know which type of vehicle you're 
dealing with. 

Just as we can tell between cars, bikes, and unicycles by counting the number of wheels, scientists can tell 
between different elements (remember, an element is a specific type of atom) by counting the number 
of protons. If a vehicle has only one wheel, we know it's a unicycle. If an atom has only one proton, we 
know it's a hydrogen atom or, said differently, it's an atom of the element hydrogen. Similarly, a vehicle 
with two wheels is always a bike, just like an atom with two protons is always a helium atom, or an atom 
of the element helium. When we count four wheels on a vehicle, we know it's a car, and when scientists 
count four protons in an atom, they know it's a beryllium atom, or an atom of the element beryllium. The 
list goes on: an atom with three protons is a lithium atom, an atom with five protons is a boron atom, an 

www.ckl2.org 136 




Unicycle 





Bikes 



Cars 

Figure 4.13: A unicycle, three examples of cars, and 2 examples of bikes. Can you think of some rule that 
might allow you to tell all unicycles, cars and bikes apart? 



137 



www.ckl2.org 



atom with six protons is a carbon atom... in fact, we have names for atoms containing everything from 1 
proton all the way up to 117 protons. So far, the maximum number of protons scientists have been able 
to pack into a single atom is 117, and thus there are 117 known elements. (On Earth, only atoms with a 
maximum of 92 protons occur naturally.) 




Figure 4.14: You can't really distinguish between sulfur and gold based on color because both are yellowish. 
You could say that gold was shiny, but then how would you tell the difference between gold and silicon? 
Each element, however, does have a unique number of protons. Sulfur has protons, silicon has protons, 
and gold has protons. 

Since an atom of one element can be distinguished from an atom of another element by the number of 
protons in its nucleus, scientists are always interested in this number, and how this number differs between 
different elements (Figure 4.14). Therefore, scientists give this number a special name and a special 
symbol. An element's atomic number (Z) is equal to the number of protons in the nuclei of any of 
its atoms. The atomic number for hydrogen is Z = 1, because every hydrogen atom has 1 proton. The 
atomic number for helium is Z = 2 because every helium atom has 2 protons. What's the atomic number 
of carbon? 

Mass Number (A) is the Sum of Protons and Neutrons 

In the last section we learned that each type of atom or element has a specific number of protons. This 
specific number was called the element's atomic number. Of course, since neutral atoms have to have one 
electron for every proton, an element's atomic number also tells you how many electrons are in a neutral 
atom of that element. For example, hydrogen has atomic number Z = 1. This means that an atom of 
hydrogen has one proton, and, if it's neutral, one electron as well. Gold, on the other hand, has atomic 
number Z = 79, which means that an atom of gold has 79 protons if it's neutral, and 79 electrons as well. 
So we know the number of protons, and we know the number of electrons, but what about the third type 
of subatomic particle? What about the number of neutrons in an atom? 

The number of neutrons in an atom isn't important for determining atomic number; in fact, it doesn't even 
tell you which type of atom (or which element) you have. The number of neutrons is important, though, if 
you want to find a quantity known as the mass number (A). The mass number of any atom is defined 
as the sum of the protons and neutrons in the atom: 

mass numberA = (number of protons) + (number of neutrons) 



www.ckl2.org 



138 



An atom's mass number is a very easy to calculate provided you know the number of protons and neutrons 
in an atom 

Example 1: 

What is the mass number of an atom that contains 3 protons and 4 neutrons? 
Solution: 

(number of protons) = 3 
(number of neutrons) = 4 

mass number, A = (number of protons) + (number of neutrons) 
mass number, A = (3) + (4) 
mass number A = 7 

Example 2: 

What is the mass number of an atom of helium that contains 2 neutrons? 
Solution: 

(number of protons) = 2 Remember that an atom of helium always has 2 protons, 
(number of neutrons) = 2 

mass number, A = (number of protons) + (number of neutrons) 
mass number, A = (2) + (2) 
mass numberA = 4 

Why do you think that the "mass number" includes protons and neutrons, but not electrons? You have 
already learned that the mass of an electron is very, very small compared to the mass of either a proton 
or a neutron (like the mass of a penny compared to the mass of a bowling ball). Counting the number of 
protons and neutrons tells scientists about the total mass of an atom, but counting the number of electrons 
would only confuse things. 

Think of it this way - you're asked to lift a box containing some bowling balls and some pennies, but 
the box has already been taped closed. Now, if you have to decide whether or not to get help lifting the 
box, which would you prefer to know, the total number of bowling balls and pennies, or the just the total 
number of bowling balls (Figure 4.15)? Suppose you were told only the number of bowling balls. If you 
knew that there were 20 bowling balls in the box, you wouldn't lift the box on your own, but if you knew 
that there was only 1, you probably would, even if that box contained 19 pennies that you didn't know 
about. On the other hand, if, instead of being told the number of bowling balls, you were told the number 
bowling balls and pennies, your decision would be more difficult. What if you were given the number 20? 
That could mean 20 bowling balls and no pennies, or it could mean 1 bowling ball and 19 pennies. In 
fact, it could even mean 20 pennies. Unfortunately, you would have no way of knowing what was meant 
by the number 20. Certainly, you wouldn't choose to lift 20 bowling balls, but lifting 20 pennies would be 
no problem. Just like you wouldn't care about the number of pennies in the box you were about to lift, 
scientists don't care about the number of electrons when they calculate the mass number. That's why the 
mass number is only the sum of the protons and neutrons in the atom. 

Isotopes Have Varying Numbers of Neutrons 

If you were reading the last section carefully, you'll already know that you can't use the number of neutrons 
in an atom to decide which type of atom (or which element) you have. Unlike the number of protons, which 

139 www.ckl2.org 




Figure 4.15: Each of the boxes above contains a total of items. If you had to choose one to lift, though, 
you'd want to know the number of bowling balls in the each box, not the total number of items in each 
box. Obviously, you'd rather lift the box with bowling balls than the box with bowling balls. 



is always the same in atoms of the same element, the number of neutrons can be different, even in atoms 
of the same element. Atoms of the same element, containing the same number of protons, but different 
numbers of neutrons are known as isotopes. Since the isotopes of any given element all contain the same 
number of protons, they have the same atomic number (for example, the atomic number of helium is always 
2). However, since the isotopes of a given element contain different numbers of neutrons, different isotopes 
have different mass numbers. The following two examples should help to clarify this point. 

Example 3: 

What is the atomic number (Z), and the mass number of an isotope of lithium containing 3 neutrons. A 
lithium atom contains 3 protons in its nucleus. 

Solution: 

atomic numberZ = (number of protons) = 3 
(number of neutrons) = 3 

mass number, A = (number of protons) + (number of neutrons) 
mass number, A = (3) + (3) 
mass number, A = 6 

Example 4: 

What is the atomic number (Z), and the mass number of an isotope of lithium containing 4 neutrons. A 
lithium atom contains 3 protons in its nucleus. 

Solution: 

atomic numberZ = (number of protons) = 3 
(number of neutrons) = 4 

mass number, A = (number of protons) + (number of neutrons) 
mass number, A = (3) + (4) 
mass number, A = 7 

Notice that because the lithium atom always has 3 protons, the atomic number for lithium is always Z = 3. 
The mass number, however, is A = 6 in the isotope with 3 neutrons, and A = 7 in the isotope with 4 
neutrons. 

In nature, only certain isotopes exist. For instance, lithium exists as an isotope with 3 neutrons, and as 
an isotope with 4 neutrons, but it doesn't exists as an isotope with 2 neutrons, or as an isotope with 5 
neutrons. Scientists can make isotopes of lithium with 2 or 5 neutrons, but they aren't very stable (they 
fall apart easily), so they don't exist outside of the laboratory. 



www.ckl2.org 



140 



Atomic Mass is a Calculated Value 

Of course, this whole discussion of isotopes brings us back to Dalton's Atomic Theory. According to 
Dalton, atoms of a given element are identical. But if atoms of a given element can have different numbers 
of neutrons, then they can have different masses as well! How did Dalton miss this? It turns out that 
elements found in nature always exist as constant uniform mixtures of their naturally occurring isotopes. 
In other words, a piece of lithium always contains both types of naturally occurring lithium (the type 
with 3 neutrons and the type with 4 neutrons). Moreover, it always contains the two in the same relative 
amounts (or "relative abundances"). In a chunk of lithium, 93% will always be lithium with 4 neutrons, 
while the remaining 7% will always be lithium with 3 neutrons. 

Unfortunately, Dalton always experimented with large chunks of an element - chunks that contained all of 
the naturally occurring isotopes of that element. As a result, when he performed his measurements, he was 
actually observing the averaged properties of all the different isotopes in the sample. Luckily, aside from 
having different masses, most other properties of different isotopes are similar. As a result, the fact that 
atoms of a given element aren't, strictly speaking, identical, isn't all that important for most chemistry 
problems. 

Knowing about the different isotopes is important, however, when it comes to calculating atomic mass. The 
atomic mass of an element is the average mass of the masses of its isotopes and their relative percentages, 
and is typically given in "atomic mass units" (u). (Remember that an "atomic mass unit" is a convenient 
unit to use when studying atoms, because a proton is almost exactly 1.0 «). You can calculate the atomic 
mass of an element provided you know the relative abundances the element's naturally occurring isotopes, 
and the masses of those different isotopes. The examples below show how this is done. 

Example 5: 

Boron has two naturally occurring isotopes. In a sample of boron, 20% of the atoms are B-10, which is an 
isotope of boron with 5 neutrons and a mass of 10 amu. The other 80% of the atoms are B-ll, which is 
an isotope of boron with 6 neutrons and a mass of 11 amu. What is the atomic mass of boron? 

Solution: 

To do this problem, we will calculate 20% of the mass of B-10, which is how much the B-10 isotope 
contributes to the "average boron atom." We will also calculate 80% of the mass of B-ll, which is how 
much the B-ll isotope contributes to the "average boron atom." 

Step One: Convert the percentages given in the question into their decimal forms by dividing each by 100: 

20 
Decimal form of 20% = = 0.20 

100 

80 
Decimal form of 80% = = 0.80 

Step Two: Multiple the mass of each isotope by its relative abundance (percentage) in decimal form: 

20% of the mass of B - 10 = 0.20 x 10 amu = 2.0 amu 
80% of the mass of B - 11 = 0.80 x 11 amu = 8.8 amu 

Step Three: Find the total mass of the "average atom" by adding together the contributions from the 
different isotopes: 

Total mass of average atom = 2.0 u + 8.8 u = 10.8 u 

The mass of an average boron atom, and thus boron's atomic mass, is 10.8 u. 

141 www.ckl2.org 



Example 6: 

Neon has three naturally occurring isotopes. In a sample of neon, 90.92% of the atoms are Ne - 20, which 
is an isotope of neon with 10 neutrons and a mass of 19.99 u. Another 0.3% of the atoms are Ne - 21, 
which is an isotope of neon with 11 neutrons and a mass of 20.99 u. The final 8.85% of the atoms are 
Ne - 22, which is an isotope of neon with 12 neutrons and a mass of 21.99 u. What is the atomic mass of 
neon? 

Solution: 

To do this problem, we will calculate 90.9% of the mass of Ne-20, which is how much Ne-20 contributes 
to the "average neon atom". We will also calculate 0.3% of the mass of Ne - 21 and 8.8% of the mass of 
Ne - 22, which are how much the Ne - 21 isotope and the Ne - 22 isotope contribute to the "average neon 
atom" respectively. 

Step One: Convert the percentages given in the question into their decimal forms by dividing each by 100: 

90 92 
Decimal form of 90.92% = — = 0.9092 

100 

Decimal form of 0.30% = — = 0.0030 

8 85 
Decimal form of 8.85% = — = 0.0885 

Step Two: Multiple the mass of each isotope by its relative abundance (percentage) in decimal form: 

90.92% of the mass of Ne - 20 = 0.909 x 20.00 = 18.18 amu 
0.3% of the mass of Ne - 21 = 0.003 x 21.00 = 0.063 amu 
8.85% of the mass of Ne - 22 = 0.088 x 22.00 = 1.93 amu 

Step Three: Find the total mass of the "average atom" by adding together the contributions from the 
different isotopes: 

Total mass of average atom = 18.18 amu + 0.06 amu + 1.93 amu = 20.17 amu 
The mass of an average neon atom, and thus neon's atomic mass, is 20.17 amu. 

Atomic Information in the Periodic Table 

Most scientists don't want to have to calculate the atomic mass of an element every time they do an 
experiment. Nor do they want to memorize the number of protons, or the atomic number, of each of 
the 117 elements that have been discovered. As a result, this information is stored in the periodic table. 
Figure 4.16 shows a periodic table that contains more detail than the periodic table you saw back in 
Chapter 1. 

Notice that each box still contains the symbol (a capital letter or a capital letter followed by a lower case 
letter) for one of the elements, but now there are two new numbers that have been added to each square, 
one number above the element's symbol, and another number below the element's symbol. 

The number above the element's symbol in each square is the element's atomic number. Just as you learned 
previously, hydrogen (symbol H) has atomic number Z = 1, helium (symbol He) has atomic number Z = 2, 
lithium (symbol Li) has atomic number Z = 3, beryllium (symbol Be) has atomic number Z = 4, boron 
(symbol B) has atomic number Z = 5, and carbon (symbol C) has atomic number Z = 6. The number 
below the element's symbol in each square is the element's atomic mass. Notice that atomic mass of boron 

www.ckl2.org 142 



1A 






Periodic Table 


















BA 


1 


2 


a 
























He 


1.01 


2A 














3A 


4A 


5A 


6A 


7A 


4.00 


3 


4 




5 


6 


7 


8 


9 


10 


LI 


Be 














B 


c 


N 


o 


F 


Ne 


6.94 


9.01 














10.1 


12.0 


14.0 


16.0 


19.0 


20.2 


11 


12 


13 


14 


15 


16 


17 


18 


Na 


Mg 














Al 


Si 


P 


S 


CI 


Ar 


23.0 


24.3 






Transition Metals 








27.0 


28.1 


31.0 


32.1 


35.5 


40.0 


19 


20 


21 


22 


23 


24 


25 


26 


27 


28 


29 


30 


31 


32 


33 


34 


35 


36 


K 


Ca 


Sd 


Tl 


V 


Or 


Mn 


Fe 


Co 


Nl 


Ca 


Zn 


Ga 


Ge 


As 


Se 


Br 


Kr 


39.1 


40.1 


45.0 


47. 9 


£0. 9 


52.0 


54. 9 


55. 9 


58. 9 


58.7 


63. 6 


65.4 


69.7 


72. 6 


74. 9 


79.0 


79. 9 


83.8 


37 


38 


39 


40 


41 


42 


43 


44 


45 


46 


47 


48 


49 


50 


51 


52 


53 


54 


Hb 


Sr 


Y 


Zr 


Nb 


Mo 


To 


Ru 


Rh 


Pd 


Ag 


Cd 


In 


Sn 


Sb 


Te 


I 


Xe 


85.5 


87.6 


88.9 


91.2 


92.9 


95.9 


98 


101 


103 


106 


108 


112 


115 


119 


122 


128 


127 


131 


55 


56 


57 


72 


73 


74 


75 


76 


77 


78 


79 


80 


81 


82 


83 


84 


85 


86 


Cs 


Ba 


La 


Hf 


Ta 


w 


Re 


Os 


Ir 


Pt 


ATI 


Hg 


Tl 


Pb 


Bl 


Po 


At 


Bn 


133 


137 


139 


179 


181 


184 


186 


190 


192 


195 


197 


201 


204 


207 


209 


209 


210 


222 


87 


88 


89 


104 


105 




























Fr 


Ra 


Ac 


Rf 


Db 




























233 


226 


227 

































Lanthanides 



Actinides 



58 


59 


60 


61 


62 


63 


64 


65 


66 


67 


68 


69 


70 


71 


Ce 


Pt 


Nd 


Pm 


Sm 


Eu 


Gd 


Tb 


Dy 


Ho 


Er 


Tm 


Yb 


Lu 


140 


141 


144 


145 


150 


152 


157 


159 


163 


165 


167 


169 


173 


175 


90 


91 


92 


93 


94 


95 


96 


97 


98 


99 


100 


101 


102 


103 


Th 


Pa 


U 


Np 


Pu 


Am. 


Cm 


Bk 


Cf 


Es 


Fm 


Md 


No 


Lr 


232 


231 


238 


237 


244 


243 


247 


247 


251 


252 


257 


258 


259 


260 



Figure 4.16: A periodic table showing both the atomic number (Z) of each element and the mass number 
(A) of each element. 



143 



www.ckl2.org 



(symbol B) is 10.8, which is what we calculated in example 5, and the atomic mass of neon (symbol Ne) 
is 20.18, which is what we calculated in example 6. Observe how compactly the periodic table stores and 
presents a large amount of information about each element. Take time to notice that not all periodic 
tables have the atomic number above the element's symbol and the mass number below it. If you are ever 
confused, remember that the atomic number (Z) should always be the smaller of the two, while the atomic 
mass should always be the larger of the two. (The average mass must include both the number of protons 
(Z) and the average number of neutrons). 

You may listen to Tom Lehrer's humorous song "The Elements" with animation at The Element Song 
(http : //www . privatehand . com/flash/elements . html) 

Lesson Summary 

Electrons are a type of subatomic particle with a negative charge. As a result, electrons repel each 

other, but are attracted to protons. 

Protons are a type of subatomic particle with a positive charge. As a result, protons repel each other, 

but are attracted to electrons. Protons are bound together in an atom's nucleus as a result of the 

strong nuclear force. 

Neutrons are a type of subatomic particle with no charge (they're neutral). Like protons, neutrons 

are bound into the atom's nucleus as a result of the strong nuclear force. 

Protons and neutrons have approximately the same mass, but they are both much more massive than 

electrons (approximately 2,000 times as massive as an electron). 

The positive charge on a proton is equal in magnitude ("size when you ignore positive and negative 

signs") to the negative charge on an electron. As a result, a neutral atom must have an equal number 

of protons and electrons. 

Each element has a unique number of protons. An element's atomic number (Z) is equal to the 

number of protons in the nuclei of any of its atoms. 

The mass number (A) of an atom is the sum of the protons and neutrons in the atom 

mass number A = (number of protons) + (number of neutrons) 

Isotopes are atoms of the same element (same number of protons) that have different numbers of 

neutrons in their atomic nuclei. 

An element's atomic mass is the average mass of one atom of that element. An element's atomic mass 

can be calculated provided the relative abundances of the element's naturally occurring isotopes, and 

the masses of those isotopes are known. 

The periodic table is a convenient way to summarize information about the different elements. In 

addition to the element's symbol, most periodic tables will also contain the element's atomic number 

(Z), and element's atomic mass. 

Review Questions 

1. Decide whether each of the following statements is true or false. 

(a) The nucleus of an atom contains all of the protons in the atom. 

(b) The nucleus of an atom contains all of the neutrons in the atom. 

(c) The nucleus of an atom contains all of the electrons in the atom. 

(d) Neutral atoms of a given element must contain the same number of neutrons. 

(e) Neutral atoms of a given element must contain the same number of electrons. 

2. Match the subatomic property with its description. 
www.ckl2.org 144 



(a) electron - a. has an atomic charge of +le 

(b) neutron - b. has a mass of 9.109383 x 10~ 28 grams 

(c) proton - c. is neither attracted to, nor repelled from charged objects 

3. Arrange the electron, proton, and neutron in order of decreasing mass. 

4. Decide whether each of the following statements is true or false. 

(a) An element's atomic number is equal to the number of protons in the nuclei of any of its atoms. 

(b) The symbol for an element's atomic number is (A). 

(c) A neutral atom with Z = 4 must have 4 electrons. 

(d) A neutral atom with A = 4 must have 4 electrons. 

(e) An atom with 7 protons and 7 neutrons will have A = 14. 

(f) An atom with 7 protons and 7 neutrons will have Z = 14. 

(g) A neutral atom with 7 electrons and 7 neutrons will have A = 14. 

5. Use the periodic table to find the symbol for the element with: 

(a) 44 electrons in a neutral atom 

(b) 30 protons 

(c) Z = 36 

(d) an atomic mass of 14.007 amu 

6. When will the mass number (A) of an atom be... 

(a) bigger than the atomic number (Z) of the atom? 

(b) smaller than the atomic number (Z) of the atom? 

(c) equal to the atomic number (Z) of the atom? 

7. Column One contains data for 5 different elements. Column Two contains data for the same 5 
elements, however different isotopes of those elements. Match the columns by connecting isotopes of 
the same element. 

Table 4.3: 



Column One 



Column Two 



a. an atom with 2 protons and 1 
neutron 

b. a Be (beryllium) atom with 5 
neutrons 

c. an atom with Z = 6 and A = 
13 

d. an atom with 1 proton and 
A = 1 

e. an atom with Z = 7 and 7 
neutrons 



i. a C (carbon) atom with 6 neu- 
trons 

ii. an atom with 2 protons and 2 
neutrons 

iii. an atom with Z = 7 and A = 
15 

iv. an atom with A = 2 and 1 
neutron 

v. an atom with Z = 4 and 6 
neutrons 



Calculations: 

8. Match the following isotopes with their respective mass numbers. 

(a) an atom with Z = 17 and 18 neutrons - i. 35 

(b) an H atom with no neutrons - ii. 4 

(c) A He atom with 2 neutrons - iii. 1 

(d) an atom with Z = 11 and 11 neutrons - iv. 23 



145 



www.ckl2.org 



(e) an atom with 11 neutrons and 12 protons - v. 22 
9. Match the following isotopes with their respective atomic numbers. 

(a) a B (boron) atom with A = 10 - i. 8 

(b) an atom with A = 10 and 6 neutrons - ii. 2 

(c) an atom with 3 protons and 3 neutrons - iii. 3 

(d) an oxygen atom - iv. 4 

(e) an atom with A = 4 and 2 neutrons - v. 5 

10. Answer the following questions: 

(a) What's the mass number of an atom that contains 13 protons and 13 neutrons? 

(b) What's the mass number of an atom that contains 24 protons and 30 neutrons? 

11. Answer the following questions: 

(a) What's the mass number of the isotope of manganese (Mn) containing 28 neutrons? 

(b) What's the mass number of the isotope of calcium (Ca) containing 20 neutrons? 

12. Answer the following questions: 

(a) What's the atomic number of an atom that has 30 neutrons, and a mass number of A = 70? 

(b) What's the atomic number of an atom with 14 neutrons, if the mass number of the atom is 
A = 28? 

13. Answer the following questions: 

(a) What's the mass number of a neutral atom that contains 7 protons and 7 neutrons? 

(b) What's the mass number of a neutral atom that contains 7 electrons and 7 neutrons? 

(c) What's the mass number of a neutral atom that contains 5 protons, 5 electrons and 6 neutrons? 

(d) What's the mass number of a neutral atom that contains 3 electrons and 4 neutrons 

14. Answer the following questions: 

(a) What element has 32 neutrons in an atom with mass number A = 58? 

(b) What element has 10 neutrons in an atom with mass number A = 19? 

15. Copper has two naturally occurring isotopes. 69.15% of copper atoms are Cu - 63 and have a mass 
of 62.93 amu. The other 30.85% of copper atoms are Cu - 65 and have a mass of 64.93 amu. What 
is the atomic mass of copper? 

Vocabulary 

electron A type of subatomic particle with a negative charge. 

proton A type of subatomic particle with a positive charge. Protons are found in the nucleus of an atom. 

neutron A type of subatomic particle with no charge. Neutrons are found in the nucleus of an atom. 

the strong nuclear force The force that holds protons and neutrons together in the nucleus of the 
atom. The strong nuclear force is strong enough to overcome the repulsion between protons. 

atomic mass units (amu) A unit used to measure the masses of small quantities like protons, neutrons, 
electrons and atoms. It is useful, because the mass of a proton is very close to 1.0 amu. 

elementary charge (e) The magnitude of charge on one electron or one proton. You can treat elementary 
charges as a unit of charge. 

www.ckl2.org 146 



atomic number (Z) An element's atomic number is equal to the number of protons in the nuclei of any 
of its atoms. 

mass number (A) The mass number of an atom is the sum of the protons and neutrons in the atom. 



Image Sources 



(1) Richard Parsons. . CC-BY-SA. 

(2) Richard Parsons. . GNU-FLD. 

(3) Richard Parsons. . CC-BY-SA. 

(4) http : //en . wikipedia . org/wiki/Image : Christmas_Pudding . jpg. GNU-FDL. 

(5) http://en.wikipedia.Org/wiki/Image:Gachalaemerald. jpg. Public Domain, CC-BY-SA. 

(6) Richard Parsons. . CC-BY-SA. 

(7) http : //en. wikipedia. or g / wiki/ Image: Bowl ingb all . jpg. Public Domain, Public Domain. 

(8) . 

(9) http://en.wikipedia.0rg/wiki/Image:Dalton_John_desk.jpg. Public Domain. 

(10) Sharon Bewick. . CC-BY-SA. 

(11) http: //en. wikipedia. org/wiki/Image :Native_gold_nuggets. jpg 
http://en.wikipedia.0rg/wiki/Image:Silic0nCr0da. jpg. Public Domain, Public Domain, Public 
Domain. 

(12) http : //www . f lickr . com/photos/dottieday/536278579/. CC-BY-SA. 

(13) http : //en . wikipedia . org/wiki/File : Hendrik_ter_Brugghen_-_Democritus .jpg. Public 
Domain. 

(14) http : //www . cheml . com/acad/webtext/pre/chemsci . html. CC-BY-SA. 

(15) A portrait of J. J. Thomson.. Public Domain. 

(16) http: //en. wikipedia. org/wiki/File :RacingBicycle-non. JPG 
http : //en. wikipedia. or g/ wiki /Image: Roll s . arp . 850pix. jpg 
http : //en. wikipedia. org/wiki/Image : HondaCBRlOOOF . jpg 
http : //en. wikipedia. org/wiki/Image : Olpacecar . jpg 

http://www.flickr.eom/photos/76074333@N00/269798915. GNU-FDL, GNU-FDL,Public 
Domain,GNU-FDL,PublicDomain,CC-BY-SA. 



147 www.cki2.0rg 



Chapter 5 



The Bohr Model 



5.1 The Wave Form of Light 

Lesson Objectives 



The student will define the terms wavelength and frequency with respect to wave-form energy. 
The student will state the relationship between wavelength and frequency with respect to electro- 
magnetic radiation. 

The student will state the respective relationship between wavelengths and frequencies of selected 
colors on the electromagnetic spectrum. 



Introduction 

Our entire universe is made up of matter, which is anything that has mass and occupies space. You now 
know that matter is composed of small building blocks known as atoms, and that these small building 
blocks are composed of even smaller subatomic particles called protons, electrons and neutrons. Matter is 
all around you and you can use the atomic, or subatomic, description of matter to understand anything 
from the cells in your body to the planet Earth! 

Any object that you can hold or touch is matter. But our universe contains something else - something 
that you can't really touch, but that you can certainly see (in fact, you can't see without it!), and that you 
can often feel. It isn't matter, because it doesn't have any mass, nor does it occupy any space. Still it's 
fundamentally important to our everyday lives, and we most definitely have a name for it. Can you think 
of what it is? If you haven't guessed by now, the answer is light. 

Think about it for a minute - can you really talk about light using any of the ideas that we've considered so 
far in our study of matter and the universe? Light doesn't have any mass. Light doesn't occupy any space 
either. Try sticking your hand into the beam of light shining out of a flashlight. Your hand goes straight 
through as if there was nothing there! And yet there must be something there... how else can you explain 
the "brightness" that you see? If you have trouble understanding light and trying to define exactly what 
light is, you're not alone. Scientists had trouble explaining light too. In fact, we've only really understood 
light for about 100 years. 

www.ckl2.org 148 



Wave Form Energy 

The wave model of energy can be partially demonstrated with waves in a rope. Suppose we tie one end of 
a rope to a tree and hold the other end at a distance from the tree such that the rope is fully extended. 



Figure 5.1: Wave in a rope. 

If we then jerk the end of the rope up and down in a rhythmic way, the end of the rope we are holding 
goes up and down. When the piece of rope we are holding goes up and down, it pulls on the neighboring 
part of the rope which then also goes up and down. The up and down motion will be passed along to 
each succeeding part of the rope so that after a short time, the entire rope will contain a wave as shown in 
Figure 5.1. The red dotted line in the figure shows the undisturbed position of the rope before the wave 
was initiated. The humps above the undisturbed line are called crests and the dips below the undisturbed 
position are called troughs. It is important for you to recognize that the individual particles of the rope 
DO NOT move horizontally. The particles of rope only move up and down and if the wave is allowed to 
dissipate, all the particles of rope will be in exactly the same position they were in before the wave started. 
Each hump in the rope moves horizontally from the person to the tree but the particles of rope only move 
vertically. The energy that is put into the rope by jerking it up and down also moves horizontally from 
the person to the tree. The feeling that parts of the rope are moving horizontally is a visual illusion. 

If we jerk the rope up and down with a different rhythm, the wave in the rope will change its appearance in 
terms of crest height, distance between crests, and so forth, but the general shape of the wave will remain 
the same. We can characterize the wave in the rope with a few measurements. An instantaneous photo of 
the rope will freeze it so we can indicate some of the characteristic values. 

The distance from one crest to the next crest is called the wavelength of the wave (Figure 5.2). You 
could also measure the wavelength from one trough to the next or, in fact, between any two identical 
positions on successive waves. The symbol used for wavelength is the Greek letter lambda, A. The distance 
from the maximum height of a crest to the undisturbed position is called the amplitude of the wave. We 
could measure a velocity for the wave if we measure how far horizontally a crest travels in a unit of time. 
The unit for velocity would be the normal meters/second. We also need to determine a very important 
characteristic of waves called frequency. If we choose an exact position along the path of the wave and 
count how many crests pass the position per unit time, we would get a value for frequency. In everyday 
life, frequency values are often expressed as "cycles /second" or "waves/second" but when you try to use 

149 www.ckl2.org 



wavelength 



amplitude 




Figure 5.2: Characteristics of waves. 

these units in calculations, the word "cycles" or "waves" will not cancel. The proper unit for frequency has 
seconds in the denominator and "1" in the numerator. It is simple 1/s or s' 1 . This unit has been named 
"Hertz." Frequencies are often expressed in Hertz but when you are plugging numbers into mathematical 
formulas and wish to keep track of units, it is best to express frequency in units of s^ 1 . The symbol used 
for frequency is the Greek letter nu, v. Unfortunately, this Greek letter looks a very great deal like an 
italicized v. You must be very careful reading equations to be sure whether they are representing velocity, 
v, or frequency, v. To avoid this problem, this material will use a lower case / as the symbol for frequency. 
The velocity, wavelength, and frequency of a wave are related as indicated by the formula, v = fA. If the 
wavelength is expressed in meters and the frequency is expressed in s , then multiplying the wavelength 
times the frequency will yield m/s, which is the unit for velocity. 

Electromagnetic Waves 

Electromagnetic radiation (light) waves are somewhat like waves in a rope . . . except without the rope. 
Light waves do not need a medium through which to travel. They can travel through a vacuum, which is 
obvious since they come to us from the sun. The energy of an electromagnetic wave travels in a straight 
line along the path of the wave just as did the energy in a rope wave. The moving light has associated 
with it an oscillating electric field and an oscillating magnetic field. This means that along the straight 
line path of the wave, there exists a positive electric field that will reach maximum positive charge, then 
slowly collapse to zero charge, and then expand to a maximum negative charge. Along the path of the 
electromagnetic wave, this changing electric field repeats its oscillating charge over and over again. There 
is also a changing magnetic field that oscillates from maximum north pole field to maximum south pole 
field. When scientists try to draw a picture to represent this concept, they use the same picture of a wave 
that was used for rope waves and water waves. 



wavelength 



amplitude 




You should not allow yourself to think that the light travels in this weaving pattern. The light travels along 
the straight red line that represents the undisturbed position. For an electromagnetic wave, the crests and 
troughs represent oscillating fields, not the path of the light. We can still characterize light waves by their 



www.ckl2.org 



150 



wavelength, frequency, and velocity but these values will be significantly different numerically from water 
and rope waves. 

At some point along the path of the electromagnetic wave, the electric field will reach a maximum value 
(crest) and then, as the electromagnetic wave continues to move along its straight line path, the electric 
field will decrease, through zero, increase to a maximum trough, collapse back to zero again, and then 
expand to another maximum crest. We can measure along the path of the wave, the distance the wave 
travels between one crest and the succeeding crest and this distance will be the wavelength of the wave. 
The frequency of electromagnetic waves are determined in the same way as the frequency of a rope wave, 
that is, the number of full cycles that pass a point in a unit of time. The velocity of electromagnetic waves 
(in a vacuum) is the same for all waves regardless of frequency or wavelength. Every electromagnetic wave 
has a velocity of 3.00 x 10 8 m/s in a vacuum. The velocity of electromagnetic waves in air is slightly 
less than in a vacuum but so close that we will use the value for the velocity. The speed of light in a 
vacuum is symbolized by a lower case c. The relationship for the velocity, wavelength, and frequency of 
electromagnetic waves is c = Af. 

In rope waves and water waves, the amount of energy possessed by the wave is related to the amplitude 
of the wave; more energy is put into the rope if the end of the rope is jerked higher and lower. But, in 
electromagnetic radiation, the amount of energy possessed by the wave is related only to the frequency of 
the wave. In fact, the frequency of an electromagnetic wave can be converted directly to energy in Joules 
by multiplying by a conversion factor. The conversion factor is called Planck's constant and is equal to 
6.63 x 10~ 34 J • s. Sometimes, the unit for Planck's constant is given as Joules/Hertz but you can work 
that out to see the units are the same. The equation for the conversion of frequency to energy is E = hf, 
where E is the energy in Joules, h is Planck's constant in J • s, and / is the frequency in s^ 1 . 



The Electromagnetic Spectrum 

Electromagnetic waves have an extremely wide range of wavelengths, frequencies, and energies. The highest 
energy form of electromagnetic waves are gamma rays and the lowest energy form (that we have named) 
are radio waves. 



Roys 



X-Rays 



Ultraviolet 
Rays 



Shortwave 



Infrared 
Rays 


Redan 


FM 


TV 




AM 















1 x ID" 14 1 x ID -12 
Wavelength (in meters! 



x 10"* 1 x 10" 



1 x 10* 1 x 10 4 




4 x 10 



5 x 10 



6x ID" 7 
Wavelength (in meters) 



7 x 10 



High Energy 



Low Energy 



Figure 5.3: The Electromagnetic Spectrum. 



151 



www.ckl2.org 



On the far left of Figure 5.3 are the highest energy electromagnetic waves. These are called gamma 
rays and can be quite dangerous, in large numbers, to living systems. The next lower energy form of 
electromagnetic waves are called x-rays. Most of you are familiar with the penetration abilities of these 
waves. They can also be dangerous to living systems. Humans are advised to limit as much as possible the 
number of medical x-rays they have per year. Next lower, in energy, are ultraviolet rays. These rays are 
part of sunlight and the upper end of the ultraviolet range can cause sunburn and perhaps skin cancer. The 
tiny section next in the spectrum is the visible range of light . . . this section has been greatly expanded 
in the bottom half of the figure so it can be discussed in more detail. The visible range of electromagnetic 
radiation are the frequencies to which the human eye responds. Lower in the spectrum are infrared rays 
and radio waves. 

The light energies that are in the visible range are electromagnetic waves that cause the human eye to 
respond when those frequencies enter the eye. The eye sends a signal to the brain and the individual 
"sees" various colors. The highest energy waves in the visible region cause the brain to see violet and as 
the energy decreases, the colors change to blue, green, yellow, orange, and red. When the energy of the 
wave is above or below the visible range, the eye does not respond to them. When the eye receives several 
different frequencies at the same time, the colors are blended by the brain. If all frequencies of light strike 
the eye together, the brain sees white and if there are no visible frequencies striking the eye, the brain sees 
black. The objects that you see around you are light absorbers - that is, the chemicals on the surface of 
the object will absorb certain frequencies and not others. Your eyes detect the frequencies that strike your 
eye. Therefore, if your friend is wearing a red shirt, it means the dye in that shirt absorbs every frequency 
except red and the red frequencies are reflected. If your only light source was one exact frequency of blue 
light and you shined it on a shirt that was red in sunlight, the shirt would appear black because no light 
would be reflected. The light from fluorescent types of lights do not contain all the frequencies of sunlight 
and so clothes inside a store may appear to be a slightly different color than when you get them home. 

Lesson Summary 

• One model of light is that of wave-form electromagnetic radiation. 

• Light, in wave form, is characterized by its wavelength, A, frequency, /, and velocity, c. 

• The unit for wavelength is meters and the unit for frequency is either s^ 1 or Hertz. 

• The full spectrum of electromagnetic radiation has radio waves as its lowest energy, lowest frequency, 
longest wavelength end and gamma rays as its highest energy, highest frequency, shortest wavelength 
end. 

• The colors we see for an object are the blending of all the frequencies of light reflected by the object. 

Review Questions 

1. Choose the correct word in for the following statement. Blue light has a (longer or shorter) wavelength 
than red light. 

2. Choose the correct word in for the following statement. Yellow light has a (higher or lower) frequency 
than blue light. 

3. Choose the correct word in for the following statement. Green light has a (larger or smaller) energy 
than red light. 

4. If "light A" has a longer wavelength than "light B", then "light A" has "light 

B". 

A. a lower frequency than 
www.ckl2.org 152 



B. a higher frequency than 

C. the same frequency as 

5. If "light C" has a shorter wavelength than "light D", then "light C" has "light 

D". 

A. a larger energy than 

B. a smaller energy than 

C. the same energy as 

6. If "light E" has a higher frequency than "light F", then "light E" has - 

"light F". 

A. a longer wavelength than 

B. a shorter wavelength than 

C. the same wavelength as 

7. If "light G" has a higher frequency than "light H", then "light G" has - 

"light H". 

A. a larger energy than 

B. a smaller energy than 

C. the same energy as 

8. If "light J" has larger energy than "light K", then "light J" has "light 

K". 

A. a shorter wavelength than 

B. a longer wavelength than 

C. the same wavelength as 

9. Which of the following statements is true? 

A. The frequency of green light is higher than the frequency of blue light and the wavelength of green 

light is longer than the wavelength of blue light. 

B. The frequency of green light is higher than the frequency of blue light and the wavelength of green 

light is shorter than the wavelength of blue light. 

C. The frequency of green light is lower than the frequency of blue light and the wavelength of green light 

is shorter than the wavelength of blue light. 

D. The frequency of green light is lower than the frequency of blue light and the wavelength of green light 

is longer than the wavelength of blue light. 

E. The frequency of green light is the same as the frequency of blue light and the wavelength of green 

light is shorter than the wavelength of blue light. 

10. As the wavelength of electromagnetic radiation increases: 

A. its energy increases. 

B. its frequency increases. 

153 www.ckl2.org 



C. its speed increases. 

D. more than one of the above statements is true. 

E. none of the above statements is true. 

11. List three examples of electromagnetic waves. 

12. Why do white objects appear white? 

13. Name the colors present in white light in order of increasing frequency. 

14. Why do objects appear black? 

Vocabulary 

crest High point in a wave pattern (hill). 

trough Low point in a wave pattern (valley). 

amplitude of a wave The 'height' of a wave. In light waves, the amplitude is proportional to the 
brightness of the wave. 

frequency of a wave (v) The 'number' of waves passing a specific reference point per unit time. The 
frequency of a light wave determines the color of the light. 

hertz (Hz) The SI unit used to measure frequency. One Hertz is equivalent to 1 event (or one full wave 
passing by) per second. 

wavelength (A) The length of a single wave from peak to peak (or trough to trough). The wavelength of 
a light wave determines the color of the light. 

electromagnetic spectrum A list of all the possible types of light in order of decreasing frequency, or 
increasing wavelength, or decreasing energy. The electromagnetic spectrum includes gamma rays, 
X-rays, UV rays, visible light, IR radiation, microwaves and radio waves. 

5.2 The Dual Nature of Light 

Lesson Objectives 

• Explain the double-slit experiment and the photoelectric effect. 

• Explain why light is both a particle and a wave. 

• Use and understand the formula relating a light's velocity, frequency, and wavelength, c = f A 

• Use and understand the formula relating a light's frequency and energy, E = hf. 

Introduction 

Developing a theory to explain the nature of light was a difficult task. An acceptable theory in science 
is required to explain ALL the observations made on a particular phenomenon. Light, appears to have 
two different sets of behaviors under different circumstances. As you will see, sometimes light behaves 
like wave-form energy and sometimes it behaves like an extremely tiny particle. It required the very best 
scientific minds from all over the world to put together a theory to deal with the nature of light. 

www.ckl2.org 154 



The Difficulties of Denning Light Solved by Einstein and Planck 

Why was light so hard to understand? Part of the problem was that when scientists performed different 
experiments on light, they got conflicting results. Some experiments suggested that light was like a wave, 
while others suggested that light was like a particle! Let's take a look at what's meant by "like a wave" 
and "like a particle". 




Figure 5.4: Water diffracting into circular waves as it passes through the small opening between two rocks. 

Since you're probably familiar with water waves, we'll use water to explain wave behavior. Whenever a 
water wave is forced through a small opening, such as the space between the two rocks in Figure 5.4, it 
spreads out into a circular shape through a process known as diffraction. If several of these circular waves 
run into each other, they can interfere with one another and produce interesting patterns in the water. 




peaks 



troughs 



Figure 5.5: Patterns formed by colliding diffraction waves. 

Figure 5.5 shows some of these patterns. Look carefully at the red line which defines a cross-section of 
the pattern (portion (a)). That same cross-section is blown up in portion (b), where you can clearly see 
how it is composed of alternating "peaks" and "troughs." The peaks are actually extra high points in the 
waves (hills), while the troughs are extra low points (valleys). 

Imagine how surprised scientists were when they shone light through two narrow slits in a solid plate and 
saw a similar pattern of peaks (bright spots) and troughs (dark spots) on the wall opposite the plate. 
Obviously, this proved that light had some very wave-like properties. In fact, by assuming that light was 
a wave and that it diffracted through the two narrow slits in the plate, just like water waves diffract when 
they pass between rocks, scientists were even able to predict where the bright spots would occur! Figure 
5.6 shows how the results of the "double-slit" experiment could be understood in terms of light waves. 



155 



www.ckl2.org 



I 

I 

I 



I ^s bright =pots appear where 
predicted, assuming light 



travels as a wave 



Figure 5.6: Waves of light also diffract when they pass through narrow slits. 

What was also obvious from the double-slit experiment was that light could not be understood as a particle. 
Imagine rolling a particle (like a marble) through a small opening. Would you expect it to diffract? Of 
course not! You'd expect the marble to role in a straight line from the opening to the opposite wall, 
as shown in Figure 5.7. Light traveling as a particle, then, should make a single bright spot directly 
across from each slit opening. Since that wasn't what scientists observed, they knew that light couldn't be 
composed of tiny particles. 



I ! 



where bright spots would appear if 
light traveled as a particle 



where bright spots 



j, actually appear 

> where bright spots would appear if 
light traveled as a particle 



Figure 5.7: The bright spots formed when light shines through narrow slits are not straight through the 
openings. 

The wave theory of light seemed to work, at least for the double-slit experiment. Remember, though, that 
according to the scientific method, a theory should be tested with further experiments to make sure that 
it's accurate and complete. Unfortunately, the next experiment that scientists performed suggested that 
light was not a wave, but was, instead, a stream of particles! By shining light on a flat strip of metal, 
scientists found that they could knock electrons off of the metal surface. They called this phenomenon 
the photoelectric effect, and they called electrons that were bumped off photoelectrons. Why did the 
photoelectric effect prove that light wasn't a wave? The problem was that the number of photoelectrons 
produced by a beam of light didn't depend on how bright the light was, but instead depended on the light's 
color. To see why this was so important, we need to talk a little bit more about waves and light waves in 
particular. 

www.ckl2.org 156 



Suppose you were sitting on a pier looking out at the Atlantic Ocean. Which do you think would be more 
likely to knock you off the pier, a huge tidal wave or a wave like the type you might find on a calm day 
at the beach? Obviously, the tidal wave would have a better chance of knocking you off the pier, and 
that's because the tidal wave has more energy as a result of its bigger amplitude (amplitude is really just 
another name for the 'height' of the wave). The energy of a wave depends on its amplitude, and only on 
its amplitude. What does amplitude mean in terms of light waves? It turns out that in light waves, the 
amplitude is related to the brightness of the light - the brighter the light, the bigger the amplitude of the 
light wave. Now, based on what you know about tidal waves and piers, which do you think would be better 
at producing photoelectrons, a bright light, or a dim light? Naturally, you'd think that the bright light 
with its bigger amplitude light waves would have more energy and would therefore knock more electrons 
off... but that's NOT the case. 

It turns out that bright light and dim light knock exactly the same number of electrons off a strip of metal. 
What matters, instead of the brightness of the light, is the color of the light (Figure 5.8). Red light 
doesn't produce any photoelectrons, while blue light produces a lot of photoelectrons. Unlike brightness, 
which depends on the amplitude of the light waves, color depends on their frequency (for a discussion of 
frequency, see the box below). 



AAAAAAAAAAAA, 
AAAAAAAAAAAAAAAA/ 

Figure 5.8: The wavelengths of the different colors in the visible light spectrum. 

This made the photoelectric effect very puzzling to scientists because they knew that the energy of a wave 
doesn't depend on its frequency, only on its amplitude. So why did frequency matter when it came to 
photoelectrons? It was all rather mysterious. 

What is frequency? Frequency can be a difficult concept to understand, but it's really just a measure of 
how many times an event occurs in a given amount of time. In the case of waves, it's the number of waves 
that pass by a specific reference point per unit time. Figure 5.9 shows two different types of waves, one 
red and one blue. Notice how, in a single second (one full turn of the clock hand), 4 red waves pass by the 
dotted black line while 16 blue waves pass by the same reference point. We say that the blue waves have a 
higher frequency than the red waves. The SI unit used to measure frequency is the Hertz (Hz). One Hertz 
is equivalent to one event (or one full wave passing by) per second. 

How does the frequency of the light affect the length of the light waves? Take a close look at Figure 5.9 
again. What do you notice about the lengths of the blue and red waves? Obviously, the blue waves (higher 
frequency) have a shorter wavelength, while the red waves (lower frequency) have a longer wavelength. 
This has to be true, provided that the waves are traveling at the same speed. You can tell that the red 
and blue waves are traveling at the same speed, because their leading edges (marked by a red dot and a 
blue dot respectively) keep pace with each other. All light waves travel at the same speed. 

The explanation of the photoelectric effect began with a man named Max Planck. Max Planck wasn't 
actually studying the photoelectric effect himself. Instead, he was studying something known as black- 
body radiation. Black-body radiation is the light produced by a black object when you heat it up (think, 



157 www.ckl2.org 



fed Tvaxez / ^ \ 



Li ■=■■■.■!'." = 



VAMWWWWM{> otto 

■'-''Xm**'^. ***.*•* 2 red*™* /™\ 



^7 



VW^MIflMUPI WPKIIMI I* 12 Uue rains 



- - f£T) 






-© 



• 1 ™. 



Figure 5.9: Red and blue light have different wavelengths but travel at the same speed. 

for example, of a stove element that glows red when you turn it on). Like the photoelectric effect, scientists 
couldn't explain black-body radiation using the wave theory of light either. Max Planck, however, realized 
that black-body radiation could be understood by treating light like a stream of tiny energy packets (or 
particles). We now call these packets of energy "photons" or "quanta", and say that light is quantized. 

Albert Einstein applied the theory of quantized light to the photoelectric effect and found that the energy 
of the photons, or quanta of light, did depend on the light's frequency. In other words, all of a sudden 
Einstein could explain why the frequency of a beam of light and the energy of a beam of light were related. 
That made it a lot easier to understand why the number of photoelectrons produced by the light depended 
on the light's color (frequency). The only assumption that Einstein needed to make was that light was 
composed of particles. 

Wait! Sure the particle theory of light explained black-body radiation and the photoelectric effect, but 
what about the double-slit experiment? Didn't that require that light behave like a wave? Either the 
double-slit experiment was wrong, or else the photoelectric effect and black-body radiation were wrong. 
Surely, light had to be either a wave or a particle. Surely, it couldn't be both. Or could it? Albert Einstein 
suggested that maybe light wasn't exactly a wave or a particle. Maybe light was both. Albert Einstein's 
theory is known as the wave-particle duality of light, and is now fully accepted by modern scientists. 

Light Travels as a Wave 

You just learned that light can act like a particle or a wave, depending entirely on the situation. Did 
you ever play with Transformers when you were younger? If you did, maybe you can answer the following 
question: are Transformers vehicles (cars and aircraft) or are they robots? It's kind of a stupid question, 
isn't it? Obviously Transformers are both vehicles and robots. It's the same with light - light is both a 
wave and a particle. 

Even though Transformers are both vehicles and robots, when they want to get from one place to another 
quickly, they usually assume their vehicle form and use all of their vehicle properties (like wheels or airplane 
wings). Therefore, if you were trying to explain how a Transformer sped off in search of an enemy, you'd 
probably describe the Transformer in terms of its car or aircraft properties. 

Just as it's easiest to talk about Transformers traveling as vehicles, it's easiest to talk about light traveling 
as a wave. When light moves from one place to another, it uses its wave properties. That's why light 
passing through a thin slit will diffract; in the process of traveling through the slit, the light behaves like 
a wave. Keeping in mind that light travels as a wave, let's discuss some of the properties of its wave-like 

www.ckl2.org 158 



motion. 



First, and most importantly, all light waves travel, in a vacuum, at a speed of 299,792,458 m/s (or 
approximately 3.00 X 10 8 m/s). Imagine a tiny ant trying to surf by riding on top of a light wave (Figure 
5.10). Provided the ant could balance on the wave, it would move through space at 3.00 x 10 8 m/s. 



4 



Figure 5.10: An ant surfing a light wave. 

To put that number into perspective, when you go surfing at the beach, the waves you catch are moving 
at about 9 m/s. Unlike light waves, though, which all travel at exactly the same speed, ocean waves travel 
at different speeds depending on the depth of the ocean, the temperature, and even the wind! 

Previously, you learned that light can have different frequencies and different wavelengths. You also learned 
that because light always travels at the same speed 3.00 x 10 8 m/s, light waves with higher frequencies 
must have smaller wavelengths, while light waves with lower frequencies must have longer wavelengths. 
Scientists state this relationship mathematically using the formula 

c / x A 

speed of light(3.00 X 10 m/s) frequency(//z or s~ ) wavelength (m) 

where c is the speed of light, 3.00 x 10 8 m/s, / is the frequency and A is the wavelength. Remember that 
the unit we use to measure frequency is the Hertz (Hz), where 1 Hertz (Hz) is equal to 1 per second, s' 1 . 
Wavelength, since it is a distance, should be measured in the SI unit of distance, which is the meter (m). 
Let's see how the formula can be used to calculate the frequency or the wavelength of light. 

Example 1: 

What is the frequency of a purple colored light, if the purple light's wavelength is 4.45 X 10~ 7 m? 

speed of light, c = 3.00 x 10 8 m/s You always know the speed of light, even if the question doesn't give it 
to you. 

wavelength, A = 4.45 x 10~ 7 m 

c =fx A 

(3.00 x 10 8 m/s) =fx 4.45xl0 _7 m 

To solve for frequency, /, divide both sides of the equation by 4.45 x 10~ 7 m. 

3.00 x 10 8 m/s _ 4.45 x 10~ 7 m 
4.45 x 10- 7 m ~ X 4.45 x 10~ 7 m 
6.74 xlO 14 j -1 = fx (1) 

f = 6.74 x 10 14 Hz 

The frequency of the purple colored light is 6.74 X 10 14 Hz. 
Example 2: 

159 www.ckl2.org 



What is the frequency of a red colored light, if the red light's wavelength is 650 nm? 
speed of light, c = 3.00 x 10 8 m/s 

A = 650 nm 

A = (650 nm)(l X 10~ 9 m/nm) = 6.50 X 10 -7 m 

c =fx A 

(3.00 x 10 8 m/s) =fx 6.50xl0 _7 m 

To solve for frequency, /, divide both sides of the equation by 6.50 x 10~ 7 m. 

3.00 x 10 8 m/s _ 6.50 x 10~ 7 m 
6.50 x 10- 7 m ~ X 6.50 x lO" 7 m 
4.61 x 10 14 s~ l = f x (1) 

f=4.61xl0 14 Hz 

The frequency of the red colored light is 4.61 x 10 14 Hz. 

Notice that the wavelength in Example 1, 4.45 X 10~ 7 m, is smaller than the wavelength in Example 
2,6.50 X 10 -7 m, while the frequency in Example 1, 6.74 X 10 14 Hz is bigger than the frequency in Example 
2, 4.61 x 10 14 Hz. Just as you'd expect, a small wavelength corresponds to a big frequency, while a big 
wavelength corresponds to a small frequency. (If you're still not comfortable with that idea, take another 
look at Figure 6 and convince yourself of why this must be so, provided the waves travel at the same 
speed.) Let's take a look at one final example, where you have to solve for the wavelength instead of the 
frequency. 

Example 3: 

Scientists have measured the frequency of a particular light wave at 6.10 x 10 14 Hz. What is the wavelength 
of the light wave? 

speed of light, c = 3.00 x 10 8 m/s 

frequency, / = 6.10 X 10 14 Hz 

f = 6.10 x 10 s~ (To do dimensional analysis, it is easiest to change Hertz to per secondl Hz = ls~ ) 
c = f x A 
3.00 x 10 8 m/s = 6.10 x 10 14 s' 1 x A 

To solve for wavelength, A, divide both sides of the equation by 6.10 x 10 14 s^ 1 . 

3.00 x 10 8 m/s 6.10 x lO 14 ^ 1 



6.10 xlO 14 ^ 1 6.10 x 10 14 *- 1 
4.92xl0~ 7 m = (l)xl 



XA 



A = 4.92 x 10" 7 m 



The wavelength of the light is 4.92 X 10 7 m (or 492 nm, if you do the conversion). 
www.ckl2.org 160 



Light Consists of Energy Packets Called Photons 

We have already seen how the photoelectric effect proved that light wasn't completely wave-like, but 
rather, had particle-like properties too. Let's return to our comparison between Transformers and light. 
Transformers travel as vehicles; however, when Transformers battle each other, they fight as robots, not as 
cars and planes. The situation with light is similar. Light may travel as a wave, but as soon as it strikes an 
object and transfers its energy to that object, the light behaves as if it's made up of tiny energy packets, 
or particles, called photons. 

Remember, the energy of a wave depends only on the wave's amplitude, but not on the wave's frequency. 
The energy of a photon, or a light "particle", however, does depend on frequency. The relationship between 
a photon's energy and a photon's frequency is described mathematically by the formula 

E = h X / 

energy(/) Planck's constant(/z = 6.63 X 10~ / • s) frequency(//z or s~ ) 

where E is the energy of the photon, h is Planck's constant (which always has the value h = 6.63xl0 -34 J- s, 
and / is the frequency of the light. The SI unit for nergy is the Joule (J); the SI unit for frequency is the 
Hertz (or per second, s^ 1 ); the SI unit for Planck's constant is the Joule-second ( J • s). Although this 
equation came from complex mathematical models of black-body radiation, its meaning should be clear - 
the larger the frequency of the light beam, the more energy in each photon of light. 

Example 4: 

What is the energy of a photon in a stream of light with frequency 4.25 X 10 14 Hz? 

Planck's constant, 6.63 x 1(T 34 J • s 

frequency, 4.25 x 10 14 Hz 

/ = 4.25 x 10 14 s -1 (To do dimensional analysis, it is easiest to change Hertz to per second, 1 Hz = 1 s" 1 ) 

E h x / 

E = 6.63 x 10~ 34 / • s x 4.25 x lO^V 1 

E = 2.82 x KT 19 / 

The energy of a photon of light with frequency 4.25 x 10 14 s _1 is 2.82 x 10~ 19 J. 

Example 5: 

What is the frequency (in Hz) of a beam of light if each photon in the beam has energy 4.44 X 10 -22 J? 

Planck's constant, 6.63 x 10~ 34 J • s 

energy, E = 4.44 X 1CT 22 J 

E h X / 

4.44 x KT 22 ./ = 6.63 x 1(T 34 / • s x / 

To solve for frequency (f), divide both sides of the equation by 6.63 x 1CT34 J • s 

4.44 x KT 22 7 6.63 x 1(T 34 / • s 
— x f 

6.63 x 10- 34 7 • s 6.63 x lO" 34 / • s J 

6.70 xlO 11 ^ 1 = (1) x / 

/ = 6.70 x 10 ll Hz 



The frequency of the light is 6.70 X 10 Hz 



161 www.ckl2.org 



The Electromagnetic Spectrum 

When scientists speak of light in terms of its wave-like properties, they are often interested in the frequency 
and wavelength of the light. One way that we, as humans, can distinguish between light beams of different 
frequencies (and thus different wavelengths) is to use their colors. Light that is reddish colored has large 
wavelengths and small frequencies, while light that is bluish colored has small wavelengths and large 
frequencies. Humans can't, however, see all types of light. We can only see visible light. In fact, if the 
light's wavelength gets too small, the light becomes invisible to our eyes. We call this light ultraviolet 
(UV) radiation. Similarly, if the light's wavelength gets too large, the light also becomes invisible to our 
eyes. We call this light infrared (IR) radiation. 

Believe it or not, there are types of light with wavelengths even shorter than those of ultraviolet radiation. 
We call these types of light X-rays and gamma rays. We can use X-rays to create pictures of our bones, 
and gamma rays to kill bacteria in our food, but our eyes can't see either (you can see the picture that an 
X-ray makes, but you can't actually see the X-ray itself). On the other side of the spectrum, light with 
wavelengths even longer than those of infrared radiation are called microwaves and radio waves. We 
can use microwaves to heat our food, and radio waves to broadcast music, but again, our eyes can't see 
either. 



I 



wsmms 
Roys 



X-ftays 



Ultraviolet 
Rays 



Infrared 
Rays 


Rader 


FM 


TV 


Shortwave 


AM 















1 x ID" 14 1 x ID" 12 
Wavelength (in meters! 



1CT 4 1 x 10" S 



1 x 10 E 1 x 10 4 




4 x 10" 7 



High Enangy 



5 x 10 



6 x 10' 
Wavelength (in meters) 



7 x 10" 



Low Energy 



Figure 5.11: The electromagnetic spectrum. 

Scientists summarize all the possible types of light in what's known as the electromagnetic spectrum. 
Figure 5.11 shows a typical electromagnetic spectrum. As you can see, it's really just a list of all the 
possible types of light in order of increasing wavelength. Notice how visible light is right in the middle of 
the electromagnetic spectrum. Since light with a large wavelength has a small frequency and light with 
a small wavelength has a large frequency, arranging light in order of "increasing wavelength," is the same 
as arranging light in order of "decreasing frequency." This should be obvious from Figure 5.11 where, as 
you can see, wavelength increases to the right (decreases to the left), while frequency increases to the left 
(decreases to the right). 

Unlike wavelength and frequency, which are typically shown on the electromagnetic spectrum, energy is 
rarely included. You should, however, be able to predict how the energy of the light photons changes 
along the electromagnetic spectrum. Light with large frequencies contains photons with large energies, 



www.ckl2.org 



162 



while light with small frequencies contains photons with small energies. Therefore energy, like frequency, 
increases to the left (decreases to the right). 

Lesson Summary 

• When waves pass through narrow openings, they spread out into a circular shape through a process 
known as diffraction. 

When circular waves interact, they produce predictable patterns of peaks and troughs. 
When light is passed through two narrow slits, the light appears to interact in a manner similar to 
two circular waves spreading out from the slits. This suggests that light diffracts into circular waves 
when it passes through the slits and that these circular waves interact with each other. As a result, 
many scientists believed that light was wave-like. 

Shining light on a flat strip of metal knocks electrons off of the metal surface through what is known 
as the photoelectric effect. 

The number of photoelectrons produced by a beam of light depends on the color (wavelength) of the 
light but not on the brightness (amplitude) of the light. 

Since the energy of a wave should depend on the amplitude of the wave, scientists couldn't understand 
why a brighter light didn't knock more photoelectrons off of the metal. This led them to question of 
whether light was truly wave-like. 

Together Max Planck and Albert Einstein explained the photoelectric effect by assuming that light 
was actually a stream of little particles, or packets of energy known as photons or quanta. 
Scientists now believe that light is both a wave and a particle - a property which they term the 
wave-particle duality. 

Light travels as a wave. The speed of a light wave is always c = 3.00 x 10 8 m/s. The frequency, /, 
and wavelength, A, of a light wave are related by the formula c = f A. 

Light gives up its energy as a particle or photon. The energy (E) of a photon of light is related to 
the frequency, /, of the light according to the formula E = hf. 

The relationship between the frequency, the wavelength, and the energy of light are summarized in 
what's known as the electromagnetic spectrum. The electromagnetic spectrum is a list of light waves 
in order of increasing wavelength, decreasing frequency, and decreasing energy. 

Review Questions 

1. Decide whether each of the following statements is true or false: 

(a) Light always behaves like a wave. 

(b) Light always behaves like a particle. 

(c) Light travels like a particle and gives up its energy like a wave. 

(d) Light travels like a wave and gives up its energy like a particle. 

2. Which of the following experiments suggested that light was a wave, and which suggested that light 
was a particle? 

(a) the double-slit experiment 

(b) the photoelectric effect 

(c) black-body radiation 

3. Fill in each of the following blanks. 

(a) The brightness of a beam of light is determined by the of the light wave. 

(b) The color of a beam of light is determined by the of the light wave 

(frequency is also an acceptable answer) 

163 www.ckl2.org 



4. What is the name of the quantity depicted by each of the arrows in the diagram below (Figure 
5.12)? (Source: Sharon Bewick. CC-BY-SA) 



A AAA, 

VU/v 



Figure 5.12 

5. Consider light with a frequency of 4.4 x 10 14 Hz. What is the wavelength of this light? 

6. What is the frequency of light with a wavelength of 3.4 X 10 -9 m? 

7. What is the frequency of light with a wavelength of 575 nm? 

8. What is the energy of a photon in a beam of light with a frequency of 5.66 X 10 8 Hz? 

Vocabulary 

diffraction The tendency of a wave to spread out in a circular shape when passed through a small 
opening. 

double-slit experiment When light is passed through two narrowly separated openings (slits), the light 
produces a resulting pattern of peaks and troughs that suggests that light behaves like a wave. 

photoelectric effect The process whereby light shone on a metal surface knocks electrons (called pho- 
toelectrons) off of the surface of the metal. 

black-body radiation Light produced by a black object when the object is heated. 

photon or quanta of light A tiny particle-like packet of energy. 

wave- particle duality of light Einstein's theory, which concluded that light exhibits both particle and 
wave properties. 

electromagnetic spectrum A list of all the possible types of light in order of decreasing frequency, or 
increasing wavelength, or decreasing energy. The electromagnetic spectrum includes gamma rays, 
X-rays, UV rays, visible light, IR radiation, microwaves and radio waves. 

5.3 Light and the Atomic Spectra 

Lesson Objectives 

• Distinguish between continuous and discontinuous spectra. 

• Recognize that white light is actually a continuous spectrum of all possible wavelengths of light. 

• Recognize that all elements have unique atomic spectra. 

www.ckl2.org 164 



Introduction 

We now know how light can act as a wave or a particle, depending on the situation. You might wonder, 
though, why a chemistry textbook would waste a whole lesson on light. Light, like matter, is part of the 
universe, but chemists aren't responsible for studying the entire universe. Chemists are responsible for 
studying chemicals. What does light have to do with chemicals? Why do chemists need to know about 
light? 

It turns out that scientists can actually learn a lot about chemicals by observing how they interact with 
light. Different chemicals behave differently when struck with a beam of light. In fact, the same chemical 
will interact differently with differently colored beams of light. To understand what light can tell us about 
different chemicals, though, we must first look at the electromagnetic spectrum a little more carefully. 



Continuous Spectra Compared to Discontinuous Spectra 

If you have a class that requires a lot of work, you might find yourself saying something like, "I'm contin- 
uously doing homework for this class". Think about what you mean by that. You probably mean that the 
homework seems to be non-stop. Every time you finish one assignment, the teacher gives another, so you 
never get a break. When scientists use the word continuous, it has a similar meaning. It means no 
gaps, no holes, and no breaks of any kind. 

Scientists don't use the word continuous to describe homework, but they do use it to describe electro- 
magnetic spectra (spectra is just the plural word for spectrum). In the last section, you learned that an 
electromagnetic spectrum was a list of light arranged in order of increasing wavelength. A continuous 
electromagnetic spectrum, then, includes every possible wavelength of light between the wavelength at the 
beginning of the list and the wavelength at the end. If you find that definition confusing, consider the 
following example. 

Suppose you have a continuous spectrum that begins with light at a wavelength of 500 nm, and ends 
with light at a wavelength of 600 nm. Because it's continuous, that spectrum contains light with any 
wavelength between 500 nm and 600 nm. It contains light with a wavelength of 550 nm. It contains light 
with a wavelength of 545 nm. It contains light with a wavelength of 567.3 nm. It even contains light with 
a wavelength of 599.99999 nm. Write down any number (including a number with decimal places) that 
is bigger than 500 and smaller than 600. A continuous electromagnetic spectrum between 500 nm and 
600 nm will include light with a wavelength equal to the number you've written down. 

As shown in the last section, within the visible range of the electromagnetic spectrum, a light's wavelength 
corresponds to its color. Therefore, another way of defining a continuous spectrum in the visible range is 
to say that it is a spectrum which contains every possible color between the color at the beginning of the 
list and the color at the end. Figure 5.13 shows several examples of continuous spectra in the visible light 
range. The first continuous spectrum starts with a deep indigo blue and ends with red. Notice how the 
colors in this spectrum change smoothly all the way from indigo to red. There are no gaps, or missing 
colors. The same is true of the second continuous spectrum. The second spectrum again starts with a 
deep indigo blue, but this time ends with yellow. Once more the colors in the spectrum change smoothly 
without any gaps or holes, which makes the spectrum continuous. The third spectrum is also continuous, 
only this time it starts with the color green and ends with the color orange. 

Not all electromagnetic spectra are continuous. Sometimes they contain gaps or holes. Scientists call 
electromagnetic spectra that contain gaps or holes discontinuous. Let's reexamine spectra that start at 
500 nm and end at 600 nm. Discontinuous spectra in this range will include light with some, but not 
all wavelengths of light greater than 500 nm and less than 600 nm. A discontinuous spectrum might, for 
example, only contain light with wavelengths of 500 nm, 523 nm and 600 nm. Obviously you can think 

165 www.ckl2.org 




Figure 5.13: Several examples of continuous spectra in the visible range. 



of many numbers that lie between 500 and 600 which aren't included in that list (534 is one example). 
Therefore, the spectrum is discontinuous. A different discontinuous spectrum between 500 nm and 600 nm 
might contain every wavelength of light except 533 nm. In this case, almost every wavelength of light is 
included, but since 533 nm is missing, the spectrum is still discontinuous. 




Figure 5.14: Several examples of discontinuous spectra in the visible range. 

Figure 5.14 shows several examples of discontinuous spectra in the range of visible light. Again, since 
the wavelength of a beam of light corresponds to its color, you can clearly see when an electromagnetic 
spectrum in the visible range is discontinuous - there will be colors missing! In the first example, only 
a few shades of green are missing from the middle of the spectrum. Nevertheless, the missing shades of 
green make the spectrum discontinuous. The next two examples in Figure 5.14 have even bigger gaps of 
missing color, so it's even more obvious that they are discontinuous. 

The concept of a continuous spectrum compared to a discontinuous spectrum may seem a little silly. So 
what if one spectrum contains every possible wavelength, while another skips wavelengths here and there! 



www.ckl2.org 



166 



Why does that matter? To understand the importance of continuous and discontinuous spectra, we have 
to look a little more closely at the ways in which light interacts with matter, and how those interactions 
can actually produce electromagnetic spectra. 

Each Element Has Its Own Spectrum 

Light from the sun is a continuous spectrum. In other words, when you go to the beach and sunbathe, 
you are bombarded by light beams of every different wavelength in the electromagnetic spectrum. Certain 
wavelengths bounce off your skin, while others interact in ways that lead to a tan or even sun burn. For 
example, you've probably seen sunscreens that offer UV protection. Ultraviolet light has wavelengths 
that are smaller than those of visible blue light. In addition to the white light that we see, the continuous 
spectrum of light from the sun contains UV light, and light with that range of wavelengths can be dangerous 
to human skin cells. 

A sunbather trying to avoid tan lines and possibly too much UV radiation. (Source: http : //en . wikipedia . 
org/wiki/ Image : Sunbathing, jpg. Public Domain) 




Of course, when you're lying in the sun on the beach, you don't actually see a rainbow shining down on 
you, do you? Instead, you see white light. As a result, you might be skeptical and find it hard to believe 
that sunlight forms a continuous spectrum. Surely, the beams of light coming from the sun don't contain 
every possible wavelength of light... surely, they only contain the wavelengths of light corresponding to the 
'color' white. That argument might seem logical to you, but you've fallen into a common trap - white 
isn't a color. If you take a careful look at the electromagnetic spectrum, what you'll notice is that there is 
no 'white' light in the visible range. It turns out that white light does not come from light of any specific 
wavelength or range of wavelengths. Rather, in order for our eyes to see white, they must actually receive 
light of every wavelength in the entire visible spectrum. 

When sunlight passes through water, the white sunlight is spread out so that you can actually see the 
entire spectrum of brightly colored light composing it. This is what we know as a rainbow. (Source: 
http : //en . wikipedia . org/wiki/File : Surf ing_Rainbow . JPG . GNU-FDL) 

167 www.ckl2.org 




White is not a color, and there is no light with a wavelength corresponding to 'white'. Instead, white light 
is formed when light of every wavelength in the visible spectrum is mixed together. 

Remember that by passing white light through a prism, you were able to split the light into a rainbow, 
revealing all of the different colors, or different wavelengths of light that make up white light. 

Now that we know where to find continuous spectra (light from the sun or any other source of pure white 
light), let's discuss where and when discontinuous spectra appear in our world. By passing an electric 
current or an electric spark through certain types of matter, it's possible to make that matter glow. Neon 
lights are one common example of this phenomenon. When an electric current travels through neon gas, 
the neon glows bright orange. 



When an electric current is passed through neon gas, the gas glows orange, 
wikipedia . org/wiki/ Image : NeTube . jpg . CC-BY-SA) 



(Source: http://en. 




When an electric current is passed through argon gas, the gas glows blue. (Source: http : //en . wikipedia . 
org/wiki /Image: ArTube.jpg. CC-BY-SA) 




www.ckl2.org 



168 



Neon isn't the only gas that lights up when electricity passes through it. Electricity causes argon to glow 
blue and helium to glow pink. In fact, electricity causes every element in the entire periodic table to glow 
with a distinct color. As you might have guessed, the light from a glowing sample of neon or argon is 
very different from the light shining down from the sun. Unlike sunlight, which is white, elements such 
as neon and argon glow in colors, which means that the light they emit (or send forth) is missing certain 
wavelengths - if it wasn't, it would appear white to our eyes. 

Remember how sunlight spread out into a rainbow, or a continuous spectrum, when you passed it through 
a prism? Well, when you pass light from a sample of glowing hydrogen through a prism, it doesn't spread 
out into a continuous spectrum. Instead, it spreads out into a discontinuous spectrum, with only four 
lines of colored light. A similar thing happens when you pass the light from a sample of glowing neon, or 
argon, or even sodium through a prism. Instead of getting a continuous spectrum, you get a discontinuous 
spectrum composed of a series of colored lines. The particular series of colored lines that you get out of 
any specific element is called the element's atomic spectrum or emission spectrum. Each element 
has an emission spectrum that is characteristic to that element. In other words, the emission spectrum 
from sodium is always the same and is different than the emission spectrum from any other element, like 
calcium or helium, or gold. 

The emission spectrum for hydrogen. (Source: http://en.wikipedia.org/wiki/Atomic_emission_ 
spectrum. CC-BY-SA) 




The emission spectrum for sodium. (Source: http : //en . wikipedia . org/wiki/Atomic_emission_spectrum . 
CC-BY-SA) 




Emission spectra are important to scientists for two reasons. First, because an element's emission spectrum 
is characteristic of the element, scientists can often use emission spectra to determine which elements are 
present or absent in an unknown sample. If the emission spectrum from the sample contains lines of 
light that correspond to sodium's emission spectrum, then the sample contains sodium. You may have 
heard or read about scientists discussing what elements are present in some distant star, and after hearing 
that, wondered how scientists could know what elements are present in a place no one has ever been. 
Scientists determine what elements are present in distant stars by analyzing the light that comes from 
stars and finding the atomic spectrum of elements in that light. If the emission spectrum from the sample 
contains lines of light that correspond to helium's emission spectrum, then the sample contains helium. 
Second, and perhaps more importantly, the existence of atomic spectra and the fact that atomic spectra 
are discontinuous, can tell us a lot about how the atoms of each element are constructed. In general, an 
element's atomic spectrum results from the interaction between the electrons and protons within an atom 
of that element. The relationship between atomic spectra and the components of the atom will be the 
topic of the next lesson. 



Lesson Summary 

• When scientists use the word continuous, they mean something with no holes, no gaps, and no breaks. 

• A continuous electromagnetic spectrum contains every wavelength between the wavelength on which 
the spectrum starts and the wavelength on which the spectrum ends. 

• A discontinuous electromagnetic spectrum is a spectrum that contains gaps, holes, or breaks in terms 



169 www.ckl2.org 



of the wavelengths that it contains. 

• Light from the sun and, in fact, any pure white light source, produces light that contains a continuous 
spectrum of wavelengths. 

• White is not a color of light itself, but rather, results when light of every other color is mixed together. 

• When an electric current or an electric spark is passed through an element, the element will give off 
a colored glow. This glow is actually composed of light from a discontinuous spectrum that is unique 
to the each and every element. 

• We call the discontinuous spectrum produced by passing an electric current through an element the 
element's atomic spectrum or emission spectrum. 

• Atomic spectra can be used to identify elements. They also tell us a lot about the nature of matter. 

Review Questions 

1. Decide whether each of the following spectra is continuous or discontinuous. (Source: Sharon Bewick. 
CC-BY-SA) 




2. Decide whether each of the following statements is true or false. 

(a) White light has a wavelength of 760 nm. 

(b) Sodium's atomic spectrum is an example of a discontinuous spectrum. 

(c) Hydrogen's atomic spectrum is an example of a continuous spectrum. 

3. Fill in each of the following blanks with the words 'must', 'may' or 'does not.' 

(a) A continuous spectrum between 300 nm and 565 nm contain light with a wavelength 

of 356 nm. 

(b) A continuous spectrum between 1000 cm and 1.500 m contain light with a wavelength 

of 1.234 m. 

(c) A discontinuous spectrum between 234 mm and 545 mm contain light with a wavelength 

of 300 mm. 

4. Choose the correct word in each of the following statements. 

(a) A continuous spectrum between 532 nm and 894 nm contains light of every wavelength (greater 
than/less than) 532 nm and (greater than/ less than) 894 nm 

(b) A discontinuous spectrum between 532 nm and 894 nm contains light of every wavelength 
between 532 nm and 894 nm (including/except for) light with a wavelength of 650 nm 

www.ckl2.org 1T0 



5. What is another name for an atomic spectrum? 

6. When an electric current is passed through neon, it glows . 

7. When an electric spark is passed through argon, it glows . 

8. When an electric current is passed through helium, it glows . 

9. Suppose you had a solution which contained both dissolved hydrogen and dissolved sodium. If an 
electric current was passed through this solution, how many lines would you see in the emission 
spectrum? 

10. LEDs, or light emitting diodes, produce light by passing an electric current through a mixture of 
different atoms (or molecules) and then using their combined emission spectra to light up a room, or 
a string of Christmas tree lights. Why are white LEDs difficult and expensive to make? 

Vocabulary 

continuous electromagnetic spectrum A spectrum that contains every possible wavelength of light 
between the wavelength at the beginning of the list and the wavelength at the end. In the visible 
range of light, it is a spectrum which contains every possible color between the color at the beginning 
of the list and the color at the end. 

discontinuous electromagnetic spectrum A spectrum that includes some, but not all of the wave- 
lengths in the specified range. In the visible spectrum there are gaps or missing colors. 

pure white light A continuous spectrum of all possible wavelengths of light 

atomic spectrum (emission spectrum) A unique, discontinuous spectrum emitted by an element 
when an electric current is passed through a sample of that element 

5.4 The Bohr Model 

Lesson Objectives 

• Define an energy level in terms of the Bohr model. 

• Find the energy of a given Bohr orbit using the equation E n = = ^ £ . 

• Discuss how the Bohr model can be used to explain atomic spectra. 

Introduction 

In the last lesson, you learned that atoms of different elements produce different atomic spectra when they 
are struck by an electric spark or an electric current. This phenomenon is really rather puzzling. Why do 
atoms emit light when they are exposed to an electric current? Why is the emitted light only at specific 
wavelengths? Why do different elements have different atomic spectra? Surely these atomic spectra must 
tell us something about the atoms that they came from, but what does it all mean? These were the types of 
questions that scientists were asking in the early 1900s when a Danish physicist named Niels Bohr (Figure 
5.15) became interested in atomic spectra and the nature of the atom. 

Bohr Used Atomic Spectra to Develop His Model 

We have learned that light travels as a wave and gives up its energy as a particle. As a result, the 
wavelength (or color) of a light is related to the light's frequency which is, in turn, related to the light's 

171 www.ckl2.org 




Figure 5.15: A photograph of Niels Bohr. 

energy. All of these relationships were summarized in the electromagnetic spectrum. Remember the 
smaller the wavelength of the light (blue end of the visible spectrum), the larger the frequency and the 
larger energy. Similarly, the larger the wavelength of the light (red end of the visible spectrum), the smaller 
the frequency and the smaller the energy. 

When Niels Bohr began thinking about the atom and atomic spectra in general, he was aware of the 
wave-particle duality of light that had just been discovered. He knew that a specific wavelength (or color) 
of light was related to the light's energy. As a result, Bohr realized that the lines of color appearing in 
an element's atomic spectrum corresponded not only to those wavelengths of light but also, to the specific 
frequencies and, more importantly, to the specific energies of light. The important question, then, was why 
the atoms were only emitting light at very specific energies. Niels Bohr realized that this unusual result 
could be explained by proposing what he termed energy levels. What are energy levels? 

Electron Energy Is Quantized 

You have learned that a rock held above the ground would release potential energy as it fell. Niels Bohr 
realized that in order for atoms to release energy, there must be a similar 'falling' process going on inside the 
atom. Since Bohr, like Rutherford, knew that the protons in the atom were bound up in the tiny nucleus, 
the obvious sub-atomic objects that could be 'falling' inside the atom were the electrons. Therefore, Bohr 
proposed a model in which electrons circled around the nucleus and, on occasion, fell closer to the nucleus, 
releasing energy in the process. According to Bohr, the energy that came out of the atoms as their electrons 
fell towards the nucleus appeared as light. This light, he argued, produced the atomic spectra that could 
be seen whenever electric current was passed through an element. 

Of course, none of Bohr's arguments thus far explain why only certain energies appear in each element's 
atomic spectrum. If you hold a rock above the ground and drop it, it will release a specific amount of 
energy. If you raise that rock higher by a tiny amount, it will release slightly more energy. If you lower 
that rock by a tiny amount, it will release slightly less energy. In fact, it would seem you can get that rock 
to release any quantity of energy that you'd like, just by raising and lowering it to different levels above 

www.ckl2.org 172 



ground. But with electrons, the situation is clearly different. With electrons, it seems as if there are only a 
few levels from which electrons can fall. At least, that's what Bohr decided, and that's why he proposed 
the existence of the atomic energy level. 

According to Bohr, the electrons in an atom were only allowed to exist at certain energy levels. The 
electrons could jump from a lower level to a higher level when they gained energy (which they could get 
from a passing electric current, electric spark, heat, or light), and they could drop from a higher level to a 
lower level when they lost energy (which they released in the form of light). Most importantly, though, the 
electrons could not exist in between the allowed energy levels. Many people have compared Bohr's energy 
levels to a set of stairs. Think about a child jumping up and down a set of stairs. The child can rest at 
any one of the stairs and stay there. If he puts energy in, he can jump up to a higher stair. If he allows 
himself to fall, he can drop to a lower stair. The child cannot, however, hover at a level in between two of 
the stairs, just as an electron cannot hover in between two of the atom's energy levels. 

Just as children cannot hover between two steps on a staircase, Bohr suggested that electrons cannot hover 
between to energy levels in the atom. (Source: http://www.flickr.com/photos/veganstraightedge/ 
327066889/. CC-BY-SA) 




Bohr developed his energy level model further using principles from physics. In physics, the energy of a 
positive charge and a negative charge depends on the distance between them. Therefore, Bohr decided 
that his energy levels must correspond to orbits, or circular paths centered around the nucleus of the 
atom. Since an electron that was trapped in one of these orbits remained a constant distance from the 
nucleus, Bohr stated that within an orbit an electron had a constant energy. Of course, the fact that only 
certain energy levels were allowed, also meant that only certain orbits were allowed. Figure 5.16 shows a 
schematic illustration of Bohr's model. In the diagram, each circle from n = 1 to n = 4 is an allowed orbit. 

Notice how the negative electron, in any given orbit, is always the same distance from the positive nucleus 
at any place on the orbit. As a result electrons within an orbit always have the same energy. When an 
electron is hit by electricity, it can gain energy and can be bumped up to a higher energy orbit further 
away from the nucleus. On the other hand, when an electron loses energy, it falls back down to a lower 
energy orbit closer to the nucleus. The electron can never exist at distances in between allowed orbits. 
Notice how this limits the number of different transitions that the electron can make. Because the electron 
can only exist in certain orbits, and thus can have only certain energies, we say that the energy of the 
electron is quantized. 

Bohr Used a Formula to Determine Allowed Energy Levels 

Although Bohr's descriptive model of atomic orbits gives a nice explanation as to why atoms should 
have discontinuous atomic spectra, it's hard to test experimentally. Luckily, using advanced physics, the 
Bohr model can be used to derive a mathematical expression for the energies of the allowed orbits in the 
hydrogen atom. As you'll see in the next section, the energies of these orbits actually determine which 

173 www.ckl2.org 




Figure 5.16: A schematic illustration of the Bohr model of the atom. 

wavelengths of light appear in hydrogen's atomic spectrum, meaning that Bohr's model can, in fact, be 
tested experimentally. The equation predicting the energies of the allowed hydrogen orbits is: 

Rydberg constant, R = 1.097 x 10 7 mr 1 

Planck's constant, h = 6.63 x 1CT 34 J • s 

speed of light, c = 3.00 x 10 8 m/s 






-Rxhxc 



Here E„ is the energy of the n orbit (in other words the energy of the 1 st , 2 , 3 , 4 , etc. orbit), R is the 
Rydberg constant for hydrogen (which always has the value R = 1.097 X 10 7 m" 1 , h is Planck's constant 
(which always has the value h = 6.63 x 10~ 34 J • s, c is the speed of light (which always has the value 
c = 3.00 X 10 8 m/s, and n is the number of the orbit you are interested m.n can have any integer value 
(no decimals!) from 1 to infinity, n = 1, 2, 3, 4 . . . infinity. As was shown in Figure 5.16, n = 1 is the orbit 
closest to the atomic nucleus, n = 2 is the next orbit out, n = 3 is the one after that, and so on. In other 
words, n increases with increasing distance from the nucleus. Let's take a look at several examples using 
this equation. 

Example 1: 

What is the energy of the 3 allowed orbit in the hydrogen atom? 

Planck's constant, 6.63 x 10~ 34 J • s 

Rydberg constant for hydrogen R = 1.097 X 10 7 m~ 1 (Notice how you do not need to be given h, R or c, 
since they are all constants) 

speed of light, 3.00 x 10 8 m/s 



www.ckl2.org 



n = 3 

174 



-R x h x c 
E — 

(1.097 x 10 7 m _1 ) x (6.63 x 10 _34 J.s) x (3.00 x 10 8 m/s) 



p 



(3)2 
(7.27 x 10- 27 J.s/m) x (3.00 x 10 8 m/s) 



9 
(2.18 xl0~ 18 J) 
9 



E„ = -2.42 x 10~ 19 J 

(Be careful not to drop the negative sign! These energies should always be negative) 

Example 2: 

What is the energy of the 2nd allowed orbit in the hydrogen atom? 

Planck's constant, h = 6.63 x 10~ 34 J • s 

Rydberg constant for hydrogen, R = 1.097 X 10 m (Notice, again, how you do not need to be given h, R 
or c, since they are all constants) 

speed of light, c = 3.00 x 10 8 m/s 

n = 2 

-fixhxc 

(1.097 x 10 7 m _1 ) x (6.63 x 10 _34 J.s) x (3.00 x 10 8 m/s) 

(7.27 x 10- 27 J.s/m) x (3.00 x 10 8 m/s) 
E n - - 

(2.18 xlO -18 J) 
E n - - 

E n = -5.45 x 10" 19 J 

(Again, be careful not to drop the negative sign!) 

At the moment, these probably seem like fairly boring calculations, but in the next section, we'll see 
how energy level calculations like the two above can actually be used to predict the atomic spectrum of 
hydrogen. When scientists first did this in the early 1900s, they were amazed at how well this simple 
equation predicted the colors of light emitted by hydrogen atoms. 

Atomic Spectra Produced by Electrons Changing Energy Levels 

Suppose you have $10 in your wallet when you go to the store, and you only have $4 when you come home. 
How much money have you spent at the store? The answer is simple - you've obviously spent $6. With 
electrons, the situation is similar. If they start with 10 units of energy, and fall down to 4 units of energy, 
they lose 6 units of energy in the process. To state this mathematically, we write: 

&Ef-*f — Ef — E( 

Change in energy from initial state, i, to final state, f Energy in final state,/ Energy in the state, i 



1T5 www.ckl2.org 



When electrons 'lose' energy, though, they lose it by giving it off in the form of light. We have calculated 
the energy of an electron in the 3 hydrogen orbit, -2.42 X 1CT 19 J, and the energy of an electron in the 
2 hydrogen orbit, -5.45 X 10~ 19 J. An electron falling from the 3 hydrogen orbit to the 2 hydrogen 
orbit begins with -2.42 x 10 -19 J, and ends with -5.45 x 10~ 19 J, thus it loses -3.03 x 10 -19 J in the process. 
This can be worked out mathematically. (Don't worry too much about the fact that this lost energy has a 
negative sign. Any time that an atom gains energy, the sign will be positive, and any time that an atom 
loses energy, the sign will be negative). 

initial state, i — 3 

final state, / = 2 



E f = E 2 = -5.45 x 1(T 19 J 
E t = E 3 = -2.42 x 10~ 19 J 

A£ , 3-»2 = £2 - £3 

AEf^f = (-5.45 x 1(T 19 J) - (-2.42 x 1(T 19 J) 



Remember, subtracting a negative number is the same as adding a positive number! 

AEj^ f = (-5.45 x 1(T 19 J) + (2.42 x 1(T 19 J) 
AEj^f = -3.03 x 10~ 19 J 



All of the energy lost when an electron falls from a higher energy orbit to a lower energy orbit is turned into 
light, thus when an electron falls from the 3 hydrogen orbit to the 2 hydrogen orbit, it emits a beam of 
light with an energy of 3.03 x 10~ 19 J. You should now be able convert this energy into a wavelength (first 
convert the energy into a frequency and then, convert that frequency into a wavelength ). The wavelength 
turns out to be 656 nm and, amazingly, if you look at hydrogen's atomic spectrum, shown in Figure ??, 
you'll notice that hydrogen clearly has a line of red light at exactly 656 nm! What's more, the green 486 nm 
line in hydrogen's atomic spectrum corresponds to an electron falling from the 4 th hydrogen orbit to the 
2 hydrogen orbit, the blue 434 nm line corresponds to an electron falling from the 5 th hydrogen orbit to 
the 2 nd hydrogen orbit, and the purple 410 nm line corresponds to an electron falling from the 6 th hydrogen 
orbit to the 2 nd hydrogen orbit. In other words, Bohr's model can be used to predict the exact wavelengths 
of the four visible lines in hydrogen's atomic spectrum. 



410 nm 



434 nm 



486 nm 



656 nm 




A short discussion of atomic spectra and some animation showing the spectra of elements you chose and an 
animation of electrons changing orbits with the absorbtion and emission of light can be viewed at Spectral 
Lines (http : //www . Colorado . edu/physics/2000/quantumzone/index . html) . 

What about all of the other transitions that could occur between the allowed hydrogen orbits? For example, 
where is the line that corresponds to an electron falling from the 4th orbit to the 3rd orbit? It turns out that 
the energies of all the other transitions don't lie within range of energies included in the visible spectrum. 
As a result, even though these transitions occur, your eyes can't see the light that's given off as a result. 
Clearly, then, the predictions from Bohr's model give a perfect fit when compared to the hydrogen atomic 
spectrum. There's only one small problem... 



www.ckl2.org 



176 



Bohr's Model Only Worked for Hydrogen 

If you take a look at the periodic table, what you'll notice is that hydrogen is a very special element. 
Hydrogen is the first element, and thus, it only has one electron. It turns out that Bohr's model of the 
atom worked very well provided it was used to describe atoms with only one electron. The moment that 
the Bohr model was applied to an element with more than one electron (which, unfortunately, includes 
every element except hydrogen), the Bohr model failed miserably. 

Bohr's model failed because it treated electrons according to the laws of classical physics. Unfortunately, 
those laws only apply to fairly large objects. Back when Bohr was developing his model, scientists were only 
beginning to realize that the laws of classical physics didn't apply to matter as tiny as the electron. Electrons 
are actually quantum objects, meaning that they can only be described using the laws of quantum physics. 
Many of the differences between classical physics and quantum physics become particularly important when 
two or more quantum objects interact. As a result, while Bohr's model worked for hydrogen, it became 
worse and worse at predicting the atomic spectra for atoms with more and more electrons. Even helium, 
with two electrons, was something of a disaster! 

Bohr's model explained the emission spectrum of hydrogen which previously had no explanation. The 
invention of precise energy levels for the electrons in an electron cloud and the ability of the electrons to 
gain and lose energy by moving from one energy level to another offered an explanation for how atoms were 
able to emit exact frequencies of light. Bohr calculated energies for the energy levels of hydrogen atoms 
that yielded the exact frequencies found in the hydrogen spectrum. Furthermore, those same energy levels 
predicted that hydrogen atoms would also emit frequencies of light in the infrared and ultraviolet regions 
that no one had observed previously. The subsequent discovery that those exact frequencies of infrared 
and ultraviolet light were present in the hydrogen spectrum provided even greater support for the ideas in 
the Bohr model. 

One of the problems with Bohr's theory was that it was already known that when electrons were accelerated, 
they emitted radio waves. When you study physics, you will learn that acceleration applies to speeding 
up, slowing down, and traveling in a curved path. When charged particles are accelerated, they emit 
radio waves. In fact, that is how we create radio signals ... by forcing electrons to accelerated up and 
down in an antenna. Scientists were creating radio signals in this way since 1895. Since Bohr's electrons 
were supposedly traveling around the nucleus in a circular path, they MUST emit radio waves, hence lose 
energy and collapse into the nucleus. Since the electrons in the electron cloud of an atom did not emit 
radio waves, lose energy, and collapse into the nucleus, there was some immediate doubt that the electrons 
could be traveling in a curved path around the nucleus. Bohr attempted to deal with this problem by 
suggesting that the electron cloud contained a certain number of energy levels, that each energy level could 
hold only a single electron, and that in ground state, all electrons were in the lowest available energy level. 
Under these conditions, no electron could lose energy because there was no lower energy level available. 
Electrons could gain energy and go to a higher energy level and then fall back down to the now open 
energy level thus emitting energy, but once in ground state, no lower positions were open. This explained 
why electrons circling the nucleus did not emit energy and spiral into the nucleus. Bohr did not, however, 
offer an explanation for why only the exact energy levels he calculated were present, that is, what is there 
about electrons in electron clouds that produce only a specific set of energy levels. 

Another problem with Bohr's model was the predicted positions of the electrons in the electron cloud. If 
Bohr's model was correct, the hydrogen atom electron in ground state would always be the same distance 
from the nucleus. If we could take a series of photographic snapshots of a hydrogen electron cloud that 
would freeze the position of the electron so we could see exactly where it was located at different times, 
we still wouldn't know the path the electron followed to get from place to place, but we could see a few 
positions for the electron. Such an image of electron positions would show the electron could actually be 
various distances from the nucleus rather that at a constant distance. 

ITT www.ckl2.org 



A. ••>*•. B * • . 



Figure 5.17: The electrons are NOT a constant distance from the nucleus. 



If the electron circled the nucleus as suggested by Bohr's model, the electron positions would always be 
the same distance from the nucleus as shown in Figure 5.17, portion A. In reality, however, the electron 
is found at many different distances from the nucleus as in Figure 5.17, portion B. To solve all these 
discrepancies, scientists would need a completely new way of looking at not just energy but at matter as 
well. 

The development of Bohr's model of the atom is a good example of the scientific method. It shows how 
the observations of atomic spectra lead to the invention of a hypothesis about the nature of matter to 
explain the observations. The hypothesis also made predictions about spectral lines that should exist in 
the infrared and ultraviolet regions and when these observations were found to be correct, it provided even 
more supportive evidence for the theory. Of course, further observations can also provide contradictory 
evidence that will cause the downfall of the theory which also occurred with Bohr's model. Bohr's model 
was not, however, a failure. It provided the insights that allowed the next step in the development of our 
concept of the atom. 



Bohr's Model Unacceptable 

You have just seen that the Bohr model applied classical physics to electrons when electrons can only be 
described using quantum mechanics. In addition, it turns out that Bohr's description of electrons as tiny 
little objects circling the nucleus along fixed orbits is incorrect as well. Bohr was picturing an atom that 
looked very much like a small solar system. The nucleus at the center of the atom was like the sun at the 
center of the solar system, while the electrons circling the nucleus were like the planets circling the sun. 
However, in quantum mechanics, electrons are thought of more like clouds rather than planets. Rather 
than 'circling' the nucleus confined to orbits, electrons can seem to be everywhere at once, like a fog. 

The fact that Bohr's model worked as well as it did for hydrogen is actually quite remarkable! Of course, 
Bohr's fictional 'solar system atom' wasn't a random guess. Bohr had actually thought quite a bit about 
what might be going on inside the atom, and his work marked the first major step towards understanding 
where electrons are found in the atom. Therefore, despite the fact that the Bohr model wasn't entirely 
correct, Niels Bohr was awarded a Nobel Prize for his theory in 1922. It turns out that a complete 
description of the atom and atomic spectra, requires an understanding of quantum physics. Quantum 
physics describes a bizarre world that behaves according to rules which only apply to very, very small 
objects like electrons. Back when Bohr developed his model of the atom, scientists had never heard of 
quantum physics. In fact, most scientists at the beginning of the 20th century thought that everything 
could be described using classical physics. Even today, scientists still don't fully understand quantum 
physics and the world that quantum physics describes. 



www.ckl2.org 178 



Lesson Summary 

• Niels Bohr suggested that electrons in an atom were restricted to specific orbits around the atom's 
nucleus. 

• Bohr argued that an electron in a given orbit has a constant energy, thus he named these orbits 
energy levels. 

• When an electron gains energy (from an electric current or an electric spark), it can use this energy 
to jump from a lower energy orbit (closer to the nucleus) to a higher energy orbit (farther from the 
nucleus). 

• When an electron falls from a higher energy orbit (farther from the nucleus) to a lower energy orbit 
(closer to the nucleus) it releases energy in the form of light. 

• White is not a color of light itself, but rather, results when light of every other color is mixed together 

• In Bohr's model, electrons can only exist in certain orbits and thus, can only have certain energies. 
As a result, we say that the energies of the electrons are quantized. 

• Bohr used the formula E„ = = ^j^ to predict he energy level of an electron in the nth energy level (or 
orbit) of a hydrogen atom. 

• Because, the electron is only allowed to exist at certain energy levels according to the Bohr model, 
there are only a few possible energies of light which can be released when electrons fall from one 
energy level to another. As a result, the Bohr model explains why atomic spectra are discontinuous. 

• The Bohr model successfully predicts the four colored lines in hydrogen's atomic spectrum, but it 
fails miserably when applied to any atom with more than one electron. This is due to the differences 
between the laws of classical physics and the laws of quantum physics. 

• The Bohr model is no longer accepted as a valid model of the atom. 

Review Questions 

1. Decide whether each of the following statements is true or false: 

(a) Niels Bohr suggested that the electrons in an atom were restricted to specific orbits and thus 
could only have certain energies. 

(b) Bohr's model of the atom can be used to accurately predict the emission spectrum of hydrogen. 

(c) Bohr's model of the atom can be used to accurately predict the emission spectrum of neon. 

(d) According to the Bohr model, electrons have more or less energy depending on how far around 
an orbit they have traveled. 

2. According to the Bohr model, electrons in an atom can only have certain, allowable energies. As a 
result, we say that the energies of these electrons are . 

3. The Bohr model accurately predicts the emission spectra of atoms with... 

(a) less than 1 electron. 

(b) less than 2 electrons. 

(c) less than 3 electrons. 

(d) less than 4 electrons. 

4. Consider an He + atom. Like the hydrogen atom, the He + atom only contains 1 electron, and thus 
can be described by the Bohr model. Fill in the blanks in the following statements. 

(a) An electron falling from the n = 2 orbit of He + to the n = 1 orbit of He + releases - 

energy than an electron falling from the n = 3 orbit of He + to the n = 1 orbit of He + . 

(b) An electron falling from the n = 2 orbit of He + to the n = 1 orbit of He + produces light with 

a wavelength than the light produced by an electron falling from the n = 3 orbit of 

He + to the n = 1 orbit of He + . 

179 www.ckl2.org 



+ 



(c) An electron falling from the n = 2 orbit of He + to the n = 1 orbit of He produces light with a 

frequency than the light produced by an electron falling from the n = 3 orbit of He + 

to the n = 1 orbit of He + . 

5. According to the Bohr model, higher energy orbits are located (closer to/further from) the atomic 
nucleus. This makes sense since negative electrons are (attracted to/repelled from) the positive 
protons in the nucleus, meaning it must take energy to move the electrons (away from/towards) the 
nucleus of the atom. 

6. According to the Bohr model, what is the energy of an electron in the first Bohr orbit of hydrogen? 

7. According to the Bohr model, what is the energy of an electron in the tenth Bohr orbit of hydrogen? 

8. According to the Bohr model, what is the energy of an electron in the seventh Bohr orbit of hydrogen? 

9. If an electron in a hydrogen atom has an energy of -6.06 X 10~ 20 J, which Bohr orbit is it in? 

10. If an electron in a hydrogen atom has an energy of -2.69 X 10~ 20 J, which Bohr orbit is it in? 

11. If an electron falls from the 5 th Bohr orbital of hydrogen to the 3 rd Bohr orbital of hydrogen, how 
much energy is released (you can give the energy as a positive number)? 

12. If an electron falls from the Q th Bohr orbital of hydrogen to the 3 Bohr orbital of hydrogen, what 
wavelength of light is emitted? Is this in the visible light range? 

Vocabulary 

Bohr energy level Distinct energies corresponding to the orbits (or circular paths) of electrons around 
the atomic nucleus, according to Bohr's model of the atom 

Bohr model of the atom Bohr's explanation of why elements produced discontinuous atomic spectra 
when struck by an electric current. According to this model, electrons were restricted to specific 
orbits around the nucleus of the atom in a solar system like manner. 

classical physics The laws of physics that describe the interactions of large objects 

quantum mechanics The laws of physics that describe the interactions of very small (atomic or sub- 
atomic) objects. Also known as "wave mechanics" and "quantum physics". 



Image Sources 



(i 

(2 
(3 
(4 
(5 

(6 
(7 

(8 
(9 



Richard Parsons. . CC-BY-SA. 

Sharon Bewick. A schematic illustration of the Bohr model of the atom. CC-BY-SA. 
Sharon Bewick. An ant surfing a light wave.. CC-BY-SA. 
A photograph of Niels Bohr.. Public Domain. 
Richard Parsons. The Electromagnetic Spectrum.. CC-BY-SA. 
http : //www . f lickr . com/photos/f ramesof mind/554402976/. CC-BY-SA. 
Sharon Bewick. . CC-BY-SA. 
Sharon Bewick. . CC-BY-SA. 
http : //en . wikipedia . org/wiki/Image : Spectrum4websiteEval . svg. CC-BY-SA. 



www.ckl2.org 180 



(10) http : //en . wikipedia . org/wiki/Image : Gluehlampe_01_KMJ . jpg. GNU-FDL. 

(11) Patterns formed by colliding diffraction waves.. CC-BY-SA. 

(12) Sharon Bewick. . CC-BY-SA. 

(13) http : //en . wikipedia. org/wiki/Image : Gluehlampe_01_KMJ .jpg. GNU-FDL. 

(14) CK-12 Foundation. Wave in a rope.. CC-BY-SA. 

(15) Richard Parsons. . CC-BY-SA. 

(16) CK-12 Foundation. Characteristics of waves.. CC-BY-SA. 

(17) http : //en . wikipedia . org/wiki/Image : Spectrum4websiteEval . svg. CC-BY-SA. 



181 www.ckl2.org 



Chapter 6 

Quantum Mechanics Model of 
the Atom 



6.1 The Wave Particle Duality 

Lesson Objectives 



Explain the wave-particle duality of matter. 

Define the de Broglie relationship and give a general description of how it was derived. 

Use the de Broglie relationship to calculate the wavelength of an object given the object's mass and 

velocity. 



Introduction 

In The Bohr Model of the Atom chapter you learned about the double-slit experiment. The double-slit 
experiment proved that light passing through two closely spaced slits diffracted into circular waves, which 
then interfered with each other and made interesting patterns on the opposite wall. From the double-slit 
experiment, it was obvious that light behaved like a wave. But then along came a new set of results from 
black body radiation and the photoelectric effect. Both of these experiments could only be understood 
by assuming that light was a particle! Eventually, scientists had no choice but to accept the fact that 
light was actually both a wave and a particle - hence the wave-particle duality of light. Now, you probably 
didn't find the arguments in the previous chapter all that hard to accept. You probably thought to yourself 
"Fine. If Einstein says light is a wave and a particle, I'll believe him. He's the super-genius scientist, and 
I've never understood light anyhow." 

Now let's talk about something that you probably do understand (or at least think you do). Let's talk 
about matter. Before reading any further, answer the following question - "Does matter behave like a 
wave, or a particle?" Obviously it behaves like a particle, right? Didn't we already decide that matter was 
made up of tiny particles called atoms, and that those atoms were, themselves, made up of even tinier 
particles called electrons, protons and neutrons? Don't worry. Everything that you've learned so far is 
absolutely true. Atoms, electrons, protons and neutrons do behave like particles. But that's not the whole 
story. Atoms, electrons, protons and neutrons also behave like waves! In other words, matter is just like 
light in that it has both wave-like and particle-like properties. 



www.ckl2.org 182 



Electrons Were First Only Considered to Have Particle Proper- 
ties 

Do you remember how electrons were first discovered? It was all thanks to J. J. Thomson, with his clever 
cathode ray tube experiment. Briefly, J. J. Thomson cut a small hole into the anode of a cathode ray tube. 
This allowed cathode rays to pass through the anode and strike phosphor-coated glass on the other side. 
Since the phosphor glowed when it was struck by the cathode rays J. J. Thomson could actually visualize 
how the cathode ray traveled through the cathode ray tube. 

What J. J. Thomson noticed, of course, was that the cathode rays hit the glass of the cathode ray tube 
directly opposite the hole in the anode. In other words, the cathode rays traveled in a straight line from 
the cathode, through the anode, to the glass at the end of the cathode ray tube. That sounds pretty much 
like a particle, doesn't it? After all, if the electrons in the cathode ray were wave-like, wouldn't they have 
diffracted when they passed through the hole in the anode? In that case, a huge glowing circle should have 
appeared on the phosphor painted glass, rather than just a tiny glowing spot. 

Early experiments didn't show any evidence of electron diffraction, so most scientists believed that electrons 
were particles. It turned out, though, that electrons really do have wave-like properties. The wavelengths 
were just too small to diffract in the cathode ray experiment. 



Without diffraction, the electrons 
make the phosphor glow directly 
across from the hole in the anode. 



calhod 




phoptwr coatk>g 



cathode 



if the electrons had diffracted, they 
would have coated a Sa*ge 

glowing circle on (he phosphor. 




In addition, by experimenting with magnets, J. J. Thomson had proven that the electron had mass. Up 
until the 1920s, scientists had never observed wave-like behavior from an object with mass. In fact, you 
probably can't think of anything with mass that exhibits obvious wave properties. Baseballs don't diffract; 
bullets fired from neighboring rifles at a shooting range never form interference patterns. 

You might ask about water waves, like the type you see in the ocean. Technically speaking, the water in 

183 www.ckl2.org 



the ocean doesn't behave like a wave. Instead, it's the energy traveling through the water that behaves 
like a wave. Water may have mass, but the energy traveling through the water is 'mass-less.' So far we've 
established the fact that you never see an object with mass exhibit wave-like properties. Why, then, 
would anyone suggest that matter, even matter as small as an electron, could be described in terms of a 
wave-particle duality? 

What we know as ocean waves are actually energy waves passing through the water. (Source: http: 
//en.wikipedia. org/wiki/Image: Waves_in_pacif ica_l . jpg. GNU-FDL) 




de Broglie Proposed That Electrons are Particles and Waves 

We never see matter behaving like a wave, but in 1924 a French graduate student named Louis de Broglie 
(Figure 6.1) suggested that matter did in fact, have wave-like properties. It was a strange proposal on de 
Broglie's part, but even stranger was the fact that no one laughed at him! Instead, de Broglie was awarded 
a Nobel Prize in 1929. So how did de Broglie convince the scientists of his day to believe his theory? In 
order to argue against de Broglie's wave-particle duality of matter, scientists would have had to argue with 
Einstein - and no one argues with Einstein! 




Figure 6.1: Louis de Broglie 



Einstein is most famous for saying "mass is related to energy ". Of course, this is usually written out 
www.ckl2.org 184 



as an equation, rather than as words: 
E = m x c x c 



or 

E = m x c 2 



I 



Speed of Light 

(c = 3.00 s 10 s m/s) 



Energy in 
Joules 

Mass in kg 

(In The Science of Chemistry chapter, you were told that you didn't need to know where the equation 
E — mc 2 comes from. You still don't need to know where it comes from, but you do need to know what it 
means!) So Einstein was responsible for defining the amount of energy in any object with mass. Remember, 
though, that Einstein was also responsible for proposing the wave-particle duality of light. In section 5.1.3 
you learned that, because of the wave-particle duality of light, the energy of a wave can be related to the 
wave's frequency by the equation: 



E 
1 


= 1 




1 

Energy 


1 


f 




Planck's Constant 




(6.6 


3 1 10 J4 Js> 



Louis de Broglie looked at Einstein's first equation relating mass and energy. Then he looked at Planck's 
equation, relating energy and frequency (remember, in The Bohr Model of the Atom chapter you learned 
that frequency is a property associated with all waves, and is related to wavelength according to the 
equation c = f A). After thinking for a while, de Broglie said to himself, "If mass is related to energy, 
and energy is related to frequency, then mass must be related to frequency!" That's really a very 
logical argument. It's like saying "If I'm related to my brother, and my brother is related to my 
sister, then I must be related to my sister." Even though de Broglie's argument was logical, it led to a 
very illogical result. If mass is related to frequency, then an object with mass must also have frequency. 
In other words, an object with mass must have wave-like properties! This led de Broglie to propose the 
wave-particle duality of matter. But where were these so called matter waves? Why hadn't anybody seen 
them? 



de Broglie Derived an Equation for the Wavelength of a Particle 



Using Einstein's two equations, E = mc 2 and E = hf, along with the equation relating a wave's frequency 
and its wavelength, c = A f , de Broglie was able to derive the following relationship between the wavelength 
of an object to the object's mass: 



li ^ Planck's Constant 

X, = (6_63xl0" 3+ J.s) 

I m x c 

I I Ni 

"W avelength (m) 

X Speed of Light 

Mass (kg) (c = 3.00 x 10 s m ; s) 



185 www.ckl2.org 



If you're good at math, you may be able to derive the exact same formula yourself. When de Broglie looked 
at his equation, though, he realized that something was wrong. Why should the wavelength of an object 
like an electron, or a baseball, depend on the speed of light? Neither baseballs nor electrons travel at the 
speed of light - only light travels at the speed of light! The problem, of course, was that de Broglie had 
derived his relationship using equations that applied to light, and not to matter. Luckily, the problem was 
easy to fix. All de Broglie needed to do was replace the speed of light, c, with some general speed, v. 



A, = 

I m x v 

Wavelength (m) 



Planck's Constant 
(6.63 xlO" 34 J-s) 



1 



N, 



Velocity (m/s) 



Mass (kg) 



That way, the de Broglie relationship between mass and wavelength could be applied to any thing, traveling 
at any speed. It could be applied to light, traveling at the speed c = 3.00 x 10 8 m/s, it could be applied to 
a baseball, traveling at the speed v = 45 m/s, and it could be applied to a marathon runner traveling at 
the speed of v = 3 m/s. Let's try to find the wavelengths of a few common objects (Figure 6.2). 




Figure 6.2: The wavelength for a typical baseball is far, far too small to be detectable, even in a laboratory! 

Example 1: 

One of the fastest baseballs ever pitched traveled at a speed of 46.0 m/s. If the average baseball has a 
mass of 0.145 kg, what is the baseball's wavelength? 

Planck's constant, h = 6.63 x 10~ 34 J.s 



www.ckl2.org 



186 



(Notice that you always know Planck's constant, even if the question doesn't give it to you) 
speed, v = 46.0 m/s 
mass, m = 0.145 kg 

h (6.63 x 10~ 34 J • s) 6.63 x 10~ 34 J • s ,„ J • s 



9.94 x 10" 



m X v (0.145 kg)(46.0 m/s) 6.67 kg • m/s kg • m/s 

Dividing by a fraction is the same as multiplying by its reciprocal. This applies to units as well as numbers, 
so dividing by ( kg/m • s) is the same as multiplying by ( s/kg • m). 

^ = 9.94xl0- 35 ^^ 
kg • m 

In order to account for units, you have to substitute basic units for compound units. The definition of a 
Joule is, 17 = 1 kg • m 2 /s 2 . Substituting these basic units into the equation in place of Joules yields: 

w kg • m 2 s • s 

A = 9.94 x 10~ 35 -2— — x 



kg • m 

After cancelling units, the resultant unit is meters, the correct unit for wavelength. 

A = 9.94 x 10~ 35 m 

The wavelength for a typical stock car is far, far too small to be detectable. (Source: http://en. 
wikipedia.org/wiki/Image:NASCAR_practice. jpg. Public Domain) 




Example 2: 

Stock cars typically race at around a speed of 77.0 m/s. If the average stock car has a mass of 1312 kg, 
what is the stock car's wavelength? 

Planck's constant, h = 6.63 x 10 -34 J.s 

Again, you always know Planck's constant, even if the question doesn't give it to you) 

speed, v = 77.0 m/s 

mass, m = 1312 kg 

h (6.63 x 10~ 34 J • s) 6.63 x 10~ 34 J • s „ r _ w in _ 39 J • s 



A = = -^ — '— = — = 6.57 x 10" 



mxv (1310 kg)(77.0 m/s) 100870 kg • m/s ' kg • m/s 

18T www.cki2.0rg 



-39 kg ' m x S-S \*n.™ T i„„ .•„ i„„„j ....^u i.„ ™2/„2 

s 2 kg • m 



/I = 6.57 x 10 ^ — x When Joules is replaced with kg • m /s 



After cancelling units, the resultant unit is meters, the correct unit for wavelength. 

A = 6.57 x 1(T 39 m 

What do you notice about the wavelength of a typical baseball and the wavelength of a typical stock car? 
They're both extremely small, aren't they? In fact, even the strongest microscope in the world today can't 
see down to sizes like 9.94 x 10~ 35 m or 6.57 x 10 -39 m. Well, that explains why we've never seen a matter 
wave - matter waves are just too small. 

If matter waves are too small to see, then how did scientists prove that they exist? After all, the scientific 
method requires experimental evidence before a theory is accepted - and certainly before a scientist is 
awarded a Nobel Prize for that theory! Luckily, you can see matter waves. The only trick is to pick matter 
with waves that are big enough to see. Take another look at the two example questions that we just did. 
Notice how the only two factors influencing the wavelength of an object are the object's mass and the 
object's speed. (Of course, you also need to know the value of Planck's constant, h = 6.63 X 1CT 34 J • s, 
but since it's a constant, it doesn't change). In our two examples, the car was more massive than the 
baseball, and it was traveling at a faster speed. What do you notice about the car's wavelength compared 
to the baseball's wavelength? The car's wavelength was a lot smaller, wasn't it? 

It turns out that the larger the mass of an object, and the faster the speed at which the object is traveling, 
the smaller the object's wavelength. Similarly, the smaller the mass of an object, and the slower the speed 
at which the object is traveling, the larger the object's wavelength. This is a direct consequence of de 
Broglie's equation for the wavelength - because both mass, m, and speed, v, appear in the denominator 
(the lower part of the fraction), they must both be big if you want a small wavelength, and small if you 
want a big wavelength. 

Obviously, in order to see matter waves experimentally, it would be best to have a big wavelength. That 
means we need an object with a small mass and a slow speed. What's the smallest object that you know? 
It's an electron of course! Let's see if an electron wave is large enough to detect in a laboratory 

Example 3: 

What is the wavelength of an electron traveling at 1.25xl0 5 m/s if the mass of the electron is 9.11xl0 -31 kg? 

Planck's constant, h = 6.63 x 1CT 34 J • s 

Remember, you always know Planck's constant, even if the question doesn't give it to you 

speed, v = 1.25 x 10 5 m/s 

mass, m = 9.11 x 1CT 31 kg 

h (6.63 x 10~ 34 J • s) 6.63 x 1CT 34 J • s q J ■ s 

1 - - - 5.82 x 10" 9 



m xv (9.11 X 1CT 31 kg)(1.25 X 10 5 m/s) 1.14 X 1CT 25 kg • m/s ' kg • m/s 

n kg • m 2 s • s iii 

A = 5.82 x 10 = — x When Joules is replaced with kg • m/s . 

s 2 kg • m 

After cancelling units, the resultant unit is meters, the correct unit for wavelength. 

A = 5.82 x 10~ 9 m 

www.ckl2.org 188 



Now that wavelength is bigger! All right, it's still not huge, but it's big enough that, in the 1920s, scientists 
were able to find evidence of electron waves diffracting as electrons were forced through a thin metal film. 
But let's return to one of our initial questions - if electrons have wave-like properties, why didn't J. J. 
Thomson see electron diffraction in his cathode ray tube? The answer is simple - in that experiment, the 
hole that the electrons had to pass through was just too big to cause diffraction. Waves only diffract when 
their wavelengths are about the same size as the opening that they are forced through. Since electron 
waves are extremely small, there's no way that they will diffract unless they are forced through extremely 
small openings. As for your body when you walk through a door - well, if you're worried about your body 
diffracting, use your mass and the speed at which you walk through the door to figure out your wavelength. 
That should tell you about how narrow the door has to be to cause your body to diffract. From now on, 
you can avoid doors of that size! (If you do the calculation properly, you should get a number around 
1 x 10~ 35 and, of course, you couldn't fit through a door of that size). 

Lesson Summary 

• At first electrons were thought to behave only as particles; de Broglie stated that all matter has 
wave properties and used Einstein's E = mc 2 formula and the formula E = hf to derive the formula: 
A = h/(mv) to describe the wavelength A of an object with mass m traveling at speed v. 

• Matter waves are impossible to detect for ordinary objects, like baseballs and cars, because they 
are extremely small. The larger the mass, and the faster the speed of an object, the smaller its 
wavelength. 

• Scientists found evidence of electrons diffracting when forced through a thin metal film. 

Review Questions 

1. In the last chapter you learned that light has wave-like properties and particle- like properties. Can 
you think of anything else that might have both wave-like properties and particle-like properties? 

2. Decide whether each of the following statements is true or false. 

(a) Einstein was the first scientist to propose matter waves. 

(b) You can see baseballs diffract when you throw them. 

(c) The de Broglie's wave equation can only be applied to matter traveling at the speed of light. 

(d) Most matter waves are very small, and that is why scientists didn't realize matter had wave-like 
properties until the 1920s. 

3. Choose the correct word in each of the following statements. 

(a) The (more/less) massive an object is, the longer its wavelength is 

(b) The (faster/slower) an object is traveling, the shorter its wavelength is 

(c) A particle with a mass of 1.0 g has a (longer/ shorter) wavelength than a particle with a mass 
of 3.0 g if both are traveling at the same speed 

(d) A baseball moving at 10 m/s has a (longer/ shorter) wavelength than a baseball moving at 4 m/s 

4. Choose the correct word in each of the following statements 

(a) An electron has a (longer/shorter) wavelength than a proton if both are traveling at the same 
speed. 

(b) An electron wave has a (higher/lower) frequency than a proton wave if both particles are trav- 
eling at the same speed. 

(c) If you want to increase the wavelength of an electron, you should (slow the electron down/speed 
the electron up). 

189 www.ckl2.org 



5. Choose the correct statement from the options below. The factors that influence an object's wave- 
length are... 

(a) Only the speed of the object 

(b) Only the speed of light 

(c) The speed of light and the mass of the object 

(d) Only the mass of the object 

(e) The speed of the object and the mass of the object 

6. Choose the correct statement from the options below. 

(a) Light behaves only like a wave, and matter behaves only like a particle 

(b) Light behaves only like a wave, and matter behaves only like a wave 

(c) Light behaves only like a particle, and matter behaves only like a wave 

(d) Light behaves like a wave and like a particle, but matter only behaves like a particle 

(e) Light behaves only like a wave, but matter behaves like a wave and like a particle 

(f) Light behaves like a wave and like a particle, and matter behaves like a wave and like a particle 
as well 

7. Fill in each of the following blanks. 

(a) de Broglie used the equations and to derive an equation for 

the wavelength of a matter wave. 

(b) Scientists first saw matter waves by looking for them in . This was a good 

idea, because are small enough to have matter waves that can be observed in a 

laboratory. 

8. What is the wavelength of a 5.0 kg bowling ball that rolls down the lane at 2.0 m/s? 

9. If you walk through a door at 1.0 m/s, and you weight 120 lbs (or 54 kg), what is your wavelength? 
(This is also approximately the width of the door that would cause your body to diffract.) 

10. A car has a mass of 1250 kg. If the car's wavelength is 2.41 x 10~ 38 m, at what speed is the car 
traveling? 

11. A bobsled sliding down the run at 14.8 m/s has a wavelength of 1.79 x 10~ 37 m. What is the total 
mass of the bobsled? 

Further Reading / Supplemental Links 

• http : //www . stockcarroadrace . com/compete2008 . html 

• http : //vergil . chemistry . gatech . edu/notes/quantrev/node6 . html 

• http : //www . Colorado . edu/physics/2000/quantumzone/debroglie . html 

• http : //www . launc . tased . edu . au/ONLINE/SCIENCES/physics/debroglie . html 

• http://my.morningside.edu/slaven/Physics/uncertainty/uncertainty3.html 

• http : //physics . about . com/od/lightoptics/a/waveparticle . htm 

• http : //nobelprize . org/nobel_prizes/physics/laureates/1929/broglie-bio . html 
www.ckl2.org 190 



Vocabulary 

wave-particle duality of matter Matter exhibits both particle-like and wave-like properties. 

6.2 Schrodinger's Wave Functions 

Lesson Objectives 

• Distinguish between traveling and standing waves. 

• Explain why electrons form standing waves, and what this means in terms of their energies. 

• Define an electron wave function and electron density and relate these terms to the probability of 
finding an electron at any point in space. 

Introduction 

In the last lesson, you learned that electrons and, in fact, all objects with mass, have wave-like properties. 
It might be tempting to visualize matter waves as being just like ocean waves, or waves in a puddle, but it 
turns out that matter waves are special. Unlike ocean waves or puddle waves, matter waves are 'trapped' 
in space and, as a result, can never die out, escape, or disappear. If you think carefully, you'll realize that 
this isn't true of most other waves with which you are familiar. You can form waves in a puddle by stirring 
the puddle with a stick. When you do, what you'll notice is that the waves you create actually move from 
your stick out to the edge of the puddle, where they disappear. As long as you disturb the puddle with 
your stick, the puddle will have waves in it. But as soon as you leave the puddle alone, the surface of the 
puddle will become as calm as glass. Matter waves aren't like that. Unlike puddle waves, which eventually 
die out as they escape from the puddle, matter waves never do, because matter waves don't move. As a 
result, they are forever trapped in the matter that holds them. We'll talk more about these special matter 
waves in the next section. 

An Electron is Described as a Standing Wave 

Most of the waves that you're probably familiar with are known as traveling waves, because they travel 
or move. When you're sitting on your surfboard, trying to catch a good wave, you'll often look out to 
sea in the hopes of spotting a 'big one' (Figure 6.3). When you finally do, you know that even though 
the big wave may be quite a distance off, it will eventually arrive at your surfboard and carry you in to 
shore. This, of course, implies that ocean waves are traveling waves because they actually move through 
the water. Similarly, if you're in the stands watching the Oakland A's play ball, you might find yourself 
jumping up and cheering as 'the wave' passes through the stadium (Figure 6.4). Again, this is an example 
of a traveling wave, because it moves from fans at one end of the stadium to fans at the other. There are, 
however, special waves that stay in one spot. Scientists call these waves standing waves. 

In an earlier part of this text, a wave was described in which a rope was tied to a tree and a person jerked 
the other end of the rope up and down to create a wave in the rope. When a wave travels down a rope and 
encounters an immoveable boundary (like a tree) , the wave reflects off the boundary and travels back up 
the rope. This causes interference to occur between the wave traveling toward the tree and the reflected 
wave traveling back toward the person. If the person adjusts the rhythm of their hand just right, they can 
arrange for the crests and troughs of the wave moving toward the tree to exactly coincide with the crests 
and troughs of the reflected wave. When this occurs, the apparent horizontal motion of the wave ceases 
and the wave appears to "stand" in the same place in the rope. This is called a standing wave. In such 

191 www.ckl2.org 




Figure 6.3: When you're surfing, you wait for a good wave. Surfing is possible because ocean waves are 
traveling waves. In other words, ocean waves actually move through the water. 




Figure 6.4: Some fans getting ready for 'the wave' as it passes through the stadium. 



www.ckl2.org 



192 



a case, the crests and troughs will remain in the same places and nodes will appear between the crests and 
troughs where the rope does not appear to move at all. 



Figure 6.5 

In the standing wave shown in Figure 6.5, the positions of the crests and troughs remain in the same 
positions. The crests and troughs appear to exchange places above and below the center line of the rope. 
The flat places where the rope crosses the center axis line are called nodes (positions of zero displacement). 
These nodal positions do not change. 

Traveling waves appear to travel, and standing waves appear to stand still. 

Even though standing waves don't move themselves, they are actually composed of traveling waves that 
do. Standing waves form when two traveling waves traveling in opposite directions at the same speed 
combine or run into each other. In today's lab, you'll learn how to create standing waves in a jump rope by 
feeding traveling waves into the jump rope from opposite directions. Even though a standing wave doesn't 
move, it can still 'die out'. As soon as the traveling waves that form a standing wave disappear, so does 
the standing wave itself. You'll see this first hand in the jump rope experiment. When you stop flicking 
the jump rope, the jump rope slackens and the standing waves are gone. 

Why, then, are standing waves often associated with 'trapped' waves, or waves that never die out? The 
connection between standing waves and trapped waves isn't a misconception or a misunderstanding. It 
turns out that standing waves almost always form when traveling waves are 'trapped' in a small region 
of space. Imagine what would happen if you took a whole train of traveling waves, locked them up and 
threw them into jail. Those traveling waves would probably go crazy running around the jail cell trying 
to escape. No matter how hard they tried, though, they'd always end up hitting the jail cell walls. As a 
result, the poor waves would bounce back and forth and back and forth from one end of the jail cell to the 
other. Now, if there were several traveling waves trapped in the same jail cell at the same time, one set 
of waves would end up bouncing off of the left wall, at the same time (and speed) as another set of waves 
was bouncing off of the right wall. This, of course, is exactly what's required to set up a 'standing wave' 
(two waves traveling in opposite directions at the same speed). 

The electron waves that you learned about in the last lesson form standing waves as a result of being 
trapped inside the atom. What do you think might imprison an electron wave inside an atom? The 
answer, of course, is that electrons are trapped because they are strongly attracted to the protons in the 
nucleus. Using the laws of physics to describe the forces of attraction between electrons and protons, 
scientists can figure out the size and shape of any electron's jail cell. Amazingly, by knowing the size and 
shape of an electron's jail cell, scientists can tell you what a particular electron standing wave will look 
like. 

Frequently, rather than using words to describe an electron standing wave, scientists use what's known as an 
electron wave function. Wave functions for electrons, first developed by a man named Erwin Schrodinger, 

193 www.ckl2.org 



are mathematical expressions that describe the magnitude or 'height' of an electron standing wave at every 
point in space. Now, let's discuss electron energy, which is another important electron property that can 
be explained and predicted by electron standing waves and their associated wave functions. 



Each Wave Function has an Allowed Energy Value 

Electrons form standing waves whenever they're trapped inside an atom, and thus in order to understand 
and predict electron behavior, it's important to understand electron standing waves. One of the most 
important properties that electron standing waves can help to predict is electron energy. The energy of 
an electron in any atom depends on the size and shape of the electron's standing wave when it's trapped 
inside that atom. As a result, scientists can use the wave function, or the mathematical description of an 
electron's standing wave, to figure out how much energy that electron has. 

While wave functions are helpful in predicting the amount of energy an electron has, they are also helpful 
in predicting the amount of energy an electron is allowed to have. In any confined space, like a box, a jail 
cell, or an atom, only certain standing waves are possible. Why? In order to exist, a standing wave must 
begin at one side of the box and end at the other. Waves that either don't begin where the box begins, or 
don't end where the box ends aren't allowed. Figure 6.6 shows several allowed standing waves and several 
forbidden standing waves. Notice that if the wave doesn't 'fit' perfectly inside the box, it isn't allowed. 



allowed standing waves 



forbidden standing waves 











KXXX) KXXXD 



Figure 6.6: The standing waves on the left hand side of the figure are allowed to form in the brown box, 
because they fit perfectly inside the box. In contrast, the waves on the right hand side of the figure cannot 
form in the brown box because they do not fit perfectly inside the box. 

Now here's the really strange thing about describing electrons as standing waves. Since only certain 
standing waves will fit perfectly inside an atom, electrons trapped in that atom can only have certain 
electron wave functions with certain electron energies. In other words, the standing wave picture accounts 
for the fact that some energy values are ' allowed 1 (energy values associated with standing waves that 'fit' 
perfectly inside the atom) while others are ' forbidden' energy values (energy values associated with standing 
waves that do not 'fit' perfectly inside the atom). That's exactly what Bohr said when he developed his 
model to explain atomic spectra! Bohr said that electrons could exist at specific ''allowed'' energy levels, but 
that they couldn't exist between those energy levels. Bohr, however, did not have an explanation for why 
only certain energy levels were allowed. Remarkably, the standing wave description of electrons predicts 



www.ckl2.org 



194 



quantized electron energies just like the Bohr model! 

When we represent electrons inside an atom, quantum mechanics requires that the wave must "fit" inside 
the atom so that the wave meets itself with no overlap; that is, the "electron wave" inside the atom must 
be a standing wave. If the wave is to be arranged in the form of a circle so that it attaches to itself, the 
waves can only occur if there is a whole number of waves in the circle. 





Figure 6.7 

The standing wave on the left in Figure 6.7 exactly fits in the electron cloud and hence represents an 
"allowed" energy level whereas the standing wave on the right does not fit in the electron cloud and therefore 
is not an "allowed" energy level. There are only certain energies (frequencies) for which the wavelength fits 
exactly to form a standing wave. These are the same energy levels the Bohr model suggested but NOW 
there is a reason for why electrons may have ONLY these energies. 

Max Born and Probability Patterns 

There are very few scientists, if any, who can visualize the behavior of an electron during chemical bonding 
or chemical reactions as standing waves. When chemists are asked to describe the behavior of an electron 
during a chemical change, they do not describe the mathematical equations of quantum mechanics nor do 
they discuss standing waves. The behavior of electrons in chemical reactions is best understood in terms 
of a particle. 

Erwin Schrodinger's wave equation for matter waves is similar to known equations for other wave motions 
in nature. The equation describes how a wave associated with an electron varies in space as the electron 
moves under various forces. Schrodinger worked out the solutions of his equation for the hydrogen atom and 
the results agreed with the Bohr values for the energy levels of these atoms. Furthermore, the equation 
could be applied to more complicated atoms. It was found that Schrodinger's equation gave a correct 
description of an electron's behavior in almost all cases. In spite of the overwhelming success of the wave 
equation in describing electron energies, the very meaning of the waves was vague. 

A physicist named Max Born was able to attach some physical significance to the mathematics of quantum 
mechanics. Born used data from Schrodinger's equation to show the probability of finding the electron, as 
a particle, at the point in space for which Schrodinger's equation was solved. Born's ideas allowed chemists 
to visualize the results of the wave equation as probability patterns for electron positions. 

Suppose we had a camera with such a fast shutter speed that it could take a photo of an electron in an 
electron cloud and show it frozen in position. We could then take a thousand pictures of this electron at 
different times and find it in many different positions in the atom. We could then plot all the different 
electron positions on one picture. 

Figure 6.8 shows the result of plotting many different positions of a single electron in the electron cloud 

195 www.ckl2.org 






^ '&*££*#&>■. \ 



... % 







■%-S' 



Figure 6.8: The probability pattern for a single electron atom. 

of a hydrogen atom. One way of looking at this picture is as an indication of the probability of where you 
are likely to find the electron in this atom. You must recognize, of course, that the dots are not electrons; 
this atom has only one electron. The dots are positions where the electron can be found at different times. 
From this picture, it is clear that the electron spends more time near the nucleus than it does far away. As 
you move away from the nucleus, the probability of finding the electron becomes less and less. It is also 
important to note that there is no boundary for this electron cloud. That is, there is no distance from the 
nucleus where the probability becomes zero. 

For much of the work we will be doing with atoms, it is convenient to have a boundary for the atom. Most 
often, chemists choose some distance from the nucleus beyond which the probability of finding the electron 
becomes very low and arbitrarily draw in a boundary for the atom. Frequently, the boundary is placed 
such that 90-r or 95-f- of the probability for finding the electron is inside the boundary (Figure 6.9). 




Figure 6.9: Artificial boundaries for an atom. 



www.ckl2.org 



196 



Most of the time, we will be looking at drawings of atoms that show an outside boundary for the electron 
cloud. You should keep in mind, however, that the boundary is there for our convenience and there is 
no actual boundary on an atom; that is, the probability of finding the electron never becomes zero. This 
probability plot is very simple because it is for the first electron in an atom. As the atoms become more 
complicated (more energy levels and more electrons), the probability plots also become more complicated. 

Lesson Summary 

• There are two types of waves - traveling waves that move from one place to another, and standing 
waves that are stationary. Standing waves are formed when two traveling waves traveling in opposite 
directions at the same speed combine. Electrons in atoms form standing waves because they are 
trapped by the attractive forces that exist between their negative charges and the positive charges 
on the protons in the atom's nucleus. These attractive forces determine the shape and size of the 
electron's standing wave. 

• Mathematical expressions called wave functions are used to describe an electrons standing wave in 
an atom. The energy of an electron in any atom depends on the size and shape of the electron's 
standing wave. The wave function can be used to determine the energy of an electron when it is 
trapped inside an atom. 

• Electrons in atoms are only allowed to have certain energy levels (ie - those which correspond to 
standing waves that 'fit perfectly' inside the atom). All other electron energies are forbidden. The 
probability patterns for electrons (electron density) show the probability of finding the electron at a 
given point. 

Review Questions 

1. Choose the correct word in each of the following statements. 

(a) The (more/less) electron density at a given location within the atom the more likely you are to 
find the electron there. 

(b) If there is no electron density at a particular point in space, there is (no/a high) chance of 
finding the electron there. 

(c) The higher the probability of finding an electron in a certain spot, the (more/less) electron 
density there will be at that spot. 

2. The hydrogen ion, H + has no electrons. What is the total amount of electron density in a hydrogen 
atom? 

3. Decide whether each of the following statements is true or false. 

(a) Only certain electron standing waves are allowed in any particular atom. 

(b) Only certain electron energies are allowed in any particular atom 

4. The name for the mathematical expression used to describe an electron standing wave is - 

5. Choose the correct statement. 

(a) Einstein first developed the method of describing electron standing waves with wave functions 

(b) Planck first developed the method of describing electron standing waves with wave functions 

(c) de Broglie first developed the method of describing electron standing waves with wave functions 

(d) Schrodinger first developed the method of describing electron standing waves with wave functions 

6. Circle all of the statements below which are correct. 

(a) The wave function description of electrons predicts that electrons orbit the nucleus just like 
planets orbit the sun. 

197 www.ckl2.org 



(b) The wave function description of electrons predicts that electron energies are quantized 

(c) The Bohr model of the atom suggests that electron energies are quantized. 

7. Fill in the blanks. 

(a) Since only certain values are allowed for the energy of an electron in an atom, we say that 
electron energies are . 

(b) Allowed electron energies correspond to that fit perfectly in the atom. 

8. Forbidden electron energies correspond to electron standing waves that in the atom. 

Further Reading / Supplemental Links 

• http://www. chemistry.ohio-state.edu/betha/qm 

• http : //en . wikipedia . org/wiki/Many-worlds_interpretat ion 

• http : //plato . Stanford . edu/entries/qm-manyworlds 

• http://frombob.to/many.html 

• http://en.wikipedia.org 

Vocabulary 

traveling waves Waves that travel, or move. 

standing waves Waves that do not travel, or move. They are formed when two traveling waves, moving 
in opposite directions at the same speed run into each other and combine. 

electron wave function A mathematical expression to describe the magnitude, or 'height' of an electron 
standing wave at every point in space. 

electron density The square of the wave function for the electron, it is related to the probability of 
finding an electron at a particular point in space. 

6.3 Heisenberg's Contribution 

Lesson Objectives 

• Define the Heisenberg Uncertainty Principle. 

• Explain what the Heisenberg Uncertainty Principle means in terms of the position and momentum 
of an electron. 

• Explain why the Heisenberg Uncertainty Principle helps to justify the fact that a wave function can 
only predict the probable location of an electron and not its exact location. 

www.ckl2.org 198 



Introduction 

If you think back to the last lesson, you'll remember that scientists had a lot of difficulty understanding 
electron density and the wave function in terms of the wave properties and the particle properties of the 
electron. Max Born found a way to use the electron wave function to calculate electron density, and that 
electron density is actually equal to the probability of finding an electron at any point in space. This, 
however, leads to the question, why can't scientists predict where the electron will be with certainty. Why 
can they only predict the probability of finding an electron at any given point in space? Is there something 
wrong with the theory? Is it possible to improve the theory so that scientists can predict exactly where an 
electron is and where it's going? 



Heisenberg Proposed the Uncertainty Principle for Behavior of 
Electrons 

The Uncertainty Principle 

When scientists first suggested that the wave function was related to the probability of an electron being at 
a specific point in space, it raised a lot of questions. Most importantly, scientists wondered why the wave 
function could only predict the probability of finding an electron at a given location, and not the exact 
location where the electron actually was. Some scientists suggested that the wave function couldn't make 
exact predictions because it wasn't complete. They believed that the wave function was actually missing 
information that was necessary to describe electron behavior with 'certainty'. 

Some early scientists thought that perhaps the wave function could only predict the probability of an 
electron being at a given location because the wave function was missing information. Many scientists 
spent time looking for 'hidden variables' that controlled electron behavior just like the spin controls the 
movement of a baseball. The assumption was that if the correct 'hidden variables' could be found and 
included in the wave function, then the exact movement, and the exact location of an electron could be 
predicted, just as easily as we can predict the movement of larger objects like baseballs, cars and planets. 

Everything changed, however, in 1926 when a man named Werner Heisenberg (Figure 6.10) proposed 
what's known as the Heisenberg Uncertainty Principle. According to the Heisenberg Uncertainty 
Principle, it is impossible to measure certain properties, like momentum (speed multiplied by mass) and 
position at the same time without introducing uncertainty into the measurement. Of course, if you can't 
make accurate measurements, you can't make accurate prediction either. 

What prevents scientists from making accurate predictions about small objects like atoms and electrons? Is 
it that the machinery used to make the measurements is simply not good enough? Could scientists design 
better machines and better measurements and then be able to predict electron behavior with certainty? 
According to Werner Heisenberg, the answer is 'no - when it comes to small objects, scientists will never 
be able to make accurate measurements and accurate predictions, no matter how good their machinery 
is'. If you find that statement strange, you're not alone. Many scientists, even today, are bothered by the 
Heisenberg Uncertainty Principle - it seems as if, with improved machines, we should be able to make better 
measurements and thus better predictions! To some extent, that's true. Better machines can help us to 
make better measurements and better predictions, but according to the Heisenberg Uncertainty Principle, 
there is a fundamental limit to how much we can know and how accurately we can know it. It's as if there 
is 'something' in the universe which prevents us from being able to make absolutely, one hundred percent 
accurate measurements and, as a result, we will always be plagued by some uncertainty. For large objects, 
like baseballs and planets, the uncertainty is just too small to matter, but for tiny things, like atoms and 
electrons, the uncertainty becomes important. 

199 www.ckl2.org 




Figure 6.10: Werner Heisenberg 

Impossible to Fix Both the Position of an Electron and Its Momentum 

The Heisenberg Uncertainty Principle actually applies to a lot of different measurements, but often, scien- 
tists are concerned with two in particular - position and momentum. You probably know what position 
means, but momentum is a term that you don't hear a lot in everyday life. Momentum, p, is the quantity 
that you get when you multiply an object's mass by its speed (to be truly correct momentum is actually 
mass times velocity, but as in Chapter 1, we won't worry about the difference between speed and velocity). 

In terms of the position and momentum, the Heisenberg Uncertainty Principle is as follows: 

There is a fundamental limit to just how precisely we can measure both the position and the momentum 
of a particle at the same time. 

So how does the Heisenberg Uncertainty Principle relate to the electron and all the problems scientists have 
interpreting the electron wave function? Well, if you think about it logically, the Heisenberg Uncertainty 
Principle basically means that it's impossible to predict both exactly what the electron will do or exactly 
where the electron will be found. Suppose, for instance, that you know the electron's precise position, then 
according to the Heisenberg Uncertainty Principle, you can't know its precise momentum as well. In other 
words, when you know where the electron is, you don't know where it's going (since where it's 
going is determined by the velocity component of its moment). Suppose, on the other hand, that you know 
the electron's precise momentum. According to the Heisenberg Uncertainty Principle, you can't know its 
precise position as well. In other words, when you know where the electron is going, you don't 
know where it is. 

Obviously, there is always some uncertainty when it comes to electrons. You either don't know where they 
are, or else you don't know where they're going. As a result, any theory that claimed to predict exactly 
where the electron was, or exactly which path it would take as it traveled around inside the atom would 

www.ckl2.org 200 



go against the Heisenberg Uncertainty Principle. Luckily, the wave function description doesn't claim to 
predict the precise behavior of the electron. Instead, it only makes statements about the probability of 
finding the electron at one place or another. 

In other words, the wave function model is consistent with the Heisenberg Uncertainty Principle. Moreover, 
the Heisenberg Uncertainty Principle suggests that the electron wave function equation is as complete as it 
can possibly be. It may not entirely predict electron behavior, but that isn't because the model is wrong, 
or faulty. It's because, in our universe, there is a limit to how accurately we can actually know what tiny 
objects like electrons are doing. 

The Problem of Making Very Small Measurements 

When you use a thermometer to measure the temperature of a volume of water, you place the thermometer 
into the water and leave it there until the water and the thermometer have adjusted to the same tempera- 
ture. Almost all solids and liquids expand when they are heated and contract when they are cooled. Each 
substance, however, expands and contracts by different amounts. In the case of the mercury and glass in a 
thermometer, the mercury expands and contracts faster than the glass and so as a thermometer is heated, 
the mercury expands faster that the glass tube and the mercury runs up the tube. The glass has been 
marked (calibrated) for each temperature so that you can read the correct temperature from the markings 
on the tube. In the process of measuring the temperature of hot water, the thermometer is placed in the 
water and the thermometer absorbs heat from the water so that its temperature becomes the same as the 
water and you can read the temperature of the water from the temperature of the thermometer. 

You should see, at least theoretically, that when the thermometer absorbs heat from the water, the water is 
cooled down; that is, the temperature of the water decreases because of the heat lost to the thermometer. 
Therefore, the temperature you get when you measure the temperature is not the same temperature 
of the water that was present before you introduced the thermometer. The act of measuring the 
temperature of the water changed the temperature. When the volume of the water is reasonably 
large, the magnitude of the change caused by introducing the thermometer is not significant so we don't 
bother to consider it. What about if the volume of the water is very small? If the volume of water is 
200 mL and the thermometer absorbs 20 Joules of heat, the introduction of the thermometer might change 
the temperature of the water by approximately 0.03° C. Which is certainly not a significant change. But 
what if the volume of water whose temperature we were measuring was only 2 mL? Introducing the same 
thermometer into this small volume might change the temperature of the water 3°C, which would certainly 
be a significant change. The point is that when we measure very small things, the act of making the 
measurement may change what we are observing. 

Consider the method that humans use to see objects. We arrange for photons (quanta) of light to strike 
the object and we see the object by the directions, angles, and colors of the photons that bounce off the 
object and strike us in the eye or other light measuring instrument. If only red photons bounce back, we 
say the object is red. If no photons bounce back, we say there is no object present. Suppose for a moment 
that humans were gigantic stone creatures and we used golf balls to "see" with. That is, we would fire off 
golf balls at our surroundings and the balls would bounce off objects and come back and enter our eyes 
so we could see the object. If this were true, we could see mountains successfully and large buildings and 
trees . . . but could we see butterflies or small flowers? Obviously, the answer is no. The golf balls would 
simply knock small objects out of the way and continue on . . . they would not bounce back to our eyes. 

In the case of humans trying to look at electrons, the photons we use to see them with are of significant 
energy compared to electrons and when the photons collide with the electrons, the motion and/or position 
of the electron would be changed by the collision. Heisenberg's Uncertainty principle tells us we cannot 
be sure of both the location of the electron and the motion (path) of an electron at the same time. As a 
consequence, scientists had to give up the idea of knowing the path the electron follows inside 

201 www.ckl2.org 



an atom. 

Lesson Summary 

• The Heisenberg Uncertainty Principle states that it is impossible to measure certain pairs of properties 
like momentum (mass multiplied by velocity) and position at the same time without introducing 
uncertainty into one or both of the measurements. In other words, it is impossible to know both the 
exact momentum and the exact position of a particle at the same time. 

• The Heisenberg Uncertainty Principle suggests that the electron wave function is complete and that 
it does not predict the exact behavior of an electron because it is actually impossible to do so. 

• The uncertainty that Heisenberg spoke of is not due to the failure or inadequacies of the measuring 
equipment, but rather a fundamental limit imposed by our universe. 

Review Questions 

1. What types of things in everyday life are impossible to predict with absolute certainty? 

2. Why is it impossible to predict the future with absolute certainty? 

3. Fill in the blank. According to the Uncertainty Principle, it is impossible to know 

both an electron's and momentum at the same time. 

4. Decide whether each of the following statements is true or false: 

(a) According to the Heisenberg Uncertainty Principle, we will eventually be able to measure both 
an electron's exact position and its exact location at the same time. 

(b) The problem that we have when we try to measure an electron's position and its location at the 
same time is that our measuring equipment is not as good as it could be. 

(c) According to the Heisenberg Uncertainty Principle, we cannot know both the exact position and 
the exact location of a car at the same time. 

5. Circle the correct statement The Heisenberg Uncertainty Principle... 

(a) applies only to very small objects like protons and electrons 

(b) applies only to very big objects like cars and airplanes 

(c) applies to both very small objects like protons and electrons and very big objects like cars and 
airplanes 

Further Reading / Supplemental Links 

• http : //zebu . uoregon . edu/~imamur a/208/ j an27/hup . html 

• http : //www . aip . org/history/heisenberg/p08 . htm 

• http : //scienceworld . wolfram . com/physics/UncertaintyPrinciple . html 

• http : //plato . Stanford . edu/entries/qt-uncertainty 

• http://en.wikipedia.org/wiki 
www.ckl2.org 202 



Vocabulary 

momentum (p) The quantity you get when you multiply an object's mass by it's velocity (which as far 
as you're concerned is the same as its speed). 

Heisenberg's Uncertainty Principle Specific pairs of properties, such as momentum and position, are 
impossible to measure simultaneously without introducing some uncertainty. 

6.4 Quantum Numbers 

Lesson objectives 

• Explain the meaning of the principal quantum number, n. 

• Explain the meaning of the azimuthal quantum number, I. 

• Explain the meaning of the magnetic quantum number, m/. 

Introduction 

We've spent a lot of Chapter 6 talking about waves, and electron waves in particular. While most of us 
know what normal water waves look like, very few people have an understanding of what electron waves 
look like. In the last lesson, we talked about electron density, and how an electron wave could be thought 
of as representing the thickness or thinness of the electron density 'fog' at any point in space within the 
atom. We considered the probability pattern for the electron in a hydrogen atom. Now let's consider some 
more complicated atoms. 

Schrodinger's Equation Provides Three Integers Used to Define 
the Energy States and Orbital for an Electron 

You should remember that electrons form standing waves whenever they are trapped within an atom. You 
should also remember that only certain standing waves are allowed in any confined space because only 
certain standing waves can fit perfectly inside that space (remember, a perfect fit requires that a wave 
begin and end where its box begins and ends). In a one-dimensional box, it's easy to picture all of the 
possible waves that fit perfectly inside. In three dimensions, it's a little harder. Unfortunately when it 
comes to electrons in an atom, there's an additional complication on top of the fact that electron waves are 
three-dimensional! It turns out that the electrons in an atom aren't confined to nice square or rectangular 
boxes. Instead, they're confined to spherical boxes (which should make sense, since atoms are, after all, 
tiny little spheres). In other words, electron waves inside an atom must begin and end on the surface of 
a sphere. You may have a really good imagination, and an amazing ability to picture objects in three- 
dimensions, but for most people, trying to figure out what these spherical three-dimensional waves look 
like can be quite a challenge. 

Luckily, that's how electron wave functions can help. Electron wave functions basically describe the 
possible shapes that electron waves can take. We won't actually worry about the wave functions. Instead, 
we'll only worry about specific numbers, called quantum numbers. Quantum numbers are always part 
of an electron wave function and are extremely important when it comes to determining the shape of a 
probability pattern. 

When electron wave functions were first developed by a man named Erwin Schrodinger (Figure 6.11), his 
goal was to show how a wave-like description of the electron could be used to understand the behavior 

203 www.ckl2.org 



of an electron in a hydrogen atom. In order to do this, Schrodinger first defined the size and spherical 
shape of the hydrogen atom itself. Schrodinger then assumed an electron trapped within a hydrogen atom 
formed a standing wave that fit perfectly inside without 'spilling out' or 'doubling over' on itself (it turns 
out that in circular or spherical boxes, misfit waves don't 'spill out' so much as they 'double over', as shown 
in Figure 6.12. Finally, Schrodinger supposed that an electron wave had to be continuous (remember, 
something that is continuous has no gaps, holes or jumps). 




Figure 6.11: A photograph of Erwin Schrodinger 





This wave fits perfectly 
on the circle 



This wave, however, 
doubles over on itself 



Figure 6.12: When waves fit perfectly on a circle (or a sphere) they meet up with themselves and form a 
closed loop. When waves fit perfectly on a circle, though, they end up 'doubling over' on themselves. 

Amazingly, what Schrodinger discovered was that in order to satisfy his basic assumptions about the 
electron wave both fitting inside the atom and being continuous, certain quantities in the electron wave 
function had to be 'whole number integers'. Whenever Schrodinger assigned a whole number to these 
quantities, he ended up with a wave that fit perfectly inside the hydrogen atom. However, whenever he 
assigned a decimal number to these quantities, he ended up with a wave function that either 'doubled 
over' on itself, or was discontinuous! The quantities that had to be assigned whole numbers soon became 
known as quantum numbers. In the wave function for a hydrogen electron there are always three quantum 
numbers. The first quantum number, n, is called the principal quantum number, the second quantum 
number, {, is called the azimuthal quantum number, and the third quantum number, mi, is called the 



www.ckl2.org 



204 



magnetic quantum number. Together these three quantum numbers define the energy state and orbital of 
an electron, but we'll talk more about exactly what that means in the next section. 

n = Principal Quantum Number (Energy Level) 

The first quantum number, known as the principal quantum number, is given the symbol n. In order to 
describe a valid standing wave, n has to have integer values, but there's an additional restriction on n as 
well. The value of n must be a positive integer value (n = 1, 2, 3, . . .). In other words, n can never equal a 
negative integer. In fact, n can never even equal 0! The principal quantum number gives you two different 
clues as to what an electron wave looks like. First, it tells you how the electron density spreads out as you 
move away from the center of the atom. For electron waves with low principal numbers, like n = 1, the 
electron density is very thick right in close to the center of the atom, but then becomes rapidly thinner 
as you move out. In contrast, for electron waves with high principal quantum numbers, like n = 6, the 
electron density isn't as thick near the center of the atom, but is spread quite a bit further out. In general, 
the higher the principal quantum number, the further away from the nucleus you'll be able to detect a 
significant amount of electron density (Figure 6.13). 




When "n" is a small number, the 
electron probability is closer to 
the nucleus. 



X- 



#- 



When "n" is a larger number the 
electron probability is spread out 
further from the nucleus. 



Figure 6.13: When the principal quantum number, , is small, most of the electron density is found close 
to the nucleus of the atom. When the principal quantum number is large, though, the electron density 
spreads further out. 

Sometimes, you will hear people say that the principal quantum number determines the 'size' of the electron 
wave function. When people say this, they don't actually mean the absolute or total 'size' of the electron 
wave. What they mean is how big or small the electron wave appears to be based on where most (usually 
about 90%) of the electron density is concentrated. In an electron wave with a low principal quantum 
number, the electron density is mostly found close to the center of the atom. Even though there is a tiny 
bit of electron density at distances far away from the atom's nucleus, there's so little that you can't really 
tell it's there. As a result, electron waves with low principal quantum numbers appear small. On the 
other hand, in an electron waves with a high principal quantum number, the electron density is much more 
spread out, and thus much thicker at distances far from the center of the atom. Therefore, electron waves 
with high principal quantum numbers appear big. 

The principal quantum number also describes the total number of nodes that the electron wave contains. 
What are nodes? Nodes are places where the electron wave has absolutely no amplitude, or 'height'. Take 
a look at the one-dimensional waves in the figure below. Can you spot the nodes (portion a.)? Now take a 
look at the two-dimensional waves (portion b.). Can you spot the nodes? A three-dimensional wave with 
nodes is something like an onion. Think of how the onion has layers and how in between the different 
layers there's always an empty space or a break. If the onion was a three-dimensional wave, the breaks in 



205 



www.ckl2.org 



between the onion layers would be like the nodes. The higher the principal quantum number, n, the more 
nodes an electron wave has. 

A node is any place where a wave has zero amplitude, or zero height. Both a. and b. illustrate different 
ways of representing waves. In a. the amplitude corresponds to the height of the yellow line above (or 
below) the black axis. When the yellow wave crosses the black axis, the wave has zero amplitude and thus 
a node. In b., the amplitude of the wave corresponds to the thickness of the blue cloud. When there is no 
blue cloud, the wave has zero amplitude and thus a node. 

a ' a node 



AAy\AA 



. a node 



The principal quantum number is extremely important, not only because it tells you something about 
the 'size' of an electron wave and the number of nodes in an electron wave, but also because it tells you 
something about the energy of that wave. If you think back to Chapter 4, you'll remember that negative 
electrons like to be close to the positive nucleus because the energy of an electron is lower the closer it 
is to a positive charge. What does that mean in terms of the principal quantum number? It means that 
an electron wave with a lower principal quantum number, and electron density centered close to 
the nucleus of the atom, will have a lower energy, while an electron wave with a higher principal 
quantum number, and electron density spread further away from the nucleus of the atom, will have a 
higher energy. 

Similarly, as the number of nodes in an electron wave increases, the energy of the wave increases as well. 
Think about the jump rope experiment. Do you remember how you created standing waves in a jump 
rope? Did it take more or less energy on your part to get multiple waves to form along the rope? It should 
have taken you more energy to create more waves. That's because in the process of creating more waves, 
you also created more nodes and nodes are always associated with an increase in energy. Again, let's take 
a look at what his means in terms of the principal quantum number. An electron wave with a lower 
principal quantum number has fewer nodes, and thus will also have a lower energy. On the other 
hand, an electron wave with a higher principal quantum number has more nodes and thus will have 
a higher energy. 

Notice that as n increases, both the 'size' of the electron wave and the number of nodes it has increase as 
well. As a result, the energy of an electron wave always increases with n. 



www.ckl2.org 206 



1. The bigger value of the n, the higher the energy. 

2. The smaller the value of n, the lower the energy. 

Since the principal quantum number determines the energy of a particular electron wave, n is often thought 
of as referring to the 'energy level' of the electron. The term 'energy level' actually comes from Bohr's old 
solar system model of the atom. Thanks to Schrodinger, and his wave equation, though, we now know that 
the energy level doesn't correspond to a particular orbit around the nucleus, but rather, to a particular 
electron wave adopted by the electron when it becomes trapped inside the atom. 

The Angular Momentum Quantum Number (Sub-levels, s, p, d, 

f) 

The second quantum number, known as the azimuthal quantum number, is given the symbol I. While 
the principal quantum number told you about the 'size' of an electron wave and the number of nodes in 
an electron wave, the azimuthal quantum number tells you more about the 'shape' of an electron wave. 
In other words, the shape that the electron wave appears to have as a result of its electron density being 
'thicker' in one place than it is in another. You might be tempted to think that the shape of an electron 
wave is always spherical because the atom itself is spherical. It turns out, however, that while there are 
spherical electron waves, there are also waves that look like dumb-bells and waves that look like butterflies, 
and waves that look so crazy it's almost impossible to describe them! Some of the different possible shapes 
for electron waves are shown in the figure below. 




You might wonder about the various balloon-like shapes in the figure above. One difficulty with using 
drawings to represent electron waves is that the electron density itself is actually spread over a huge region 
of space surrounding the center of the atom. As you learned earlier, though, for many electron waves, 
almost all of the electron density is in close to the nucleus of the atom, with only a tiny bit of electron 
density further out. Over the years, scientists have developed a standard way of drawing electron waves. 
Rather than trying to account for all of the electron density in an electron wave, scientists usually just 
draw 'balloons' around the region of space that contains about 90% of the electron wave's total electron 
density. The figure below shows how scientists convert a cloud of electron density into a balloon. Even 
though there is a little bit of electron density outside of the cartoon balloons (and thus a small probability 
of finding the electron outside of its balloon), most electron behavior can be understood by ignoring the 
tiny bit of electron density that the cartoon balloon doesn't capture. 



207 



www.ckl2.org 




the electron density in an electron wave 




90% of the electron density is inside the 
two lobes drawn on the electron wave 



c*o 



these balloons represent the region of space 
which contains 90% of the electron density 
for this particular electron wave 



The exact shape of the cartoon balloon representing an electron wave is determined by the value of t. In 
other words, the dumb-bell shaped balloon in the image above has one value of £, while the butterfly- 
shaped balloon has another value of t. Of course, scientists would get pretty tired of saying things like 
'the dumb-bell shaped wave,' or 'the butterfly shaped wave,' so instead, they name the different waves 
using letters from the alphabet. The most common shapes for waves are called s,p,d and /. (You would 
probably find it a lot easier to remember what the different waves looked like if scientists had given them 
nice descriptive names, like the 'dumb-bell wave' or the 'butterfly wave,' but boring names like s,p,d and 
/ turn out to be much more convenient). 

In the next lesson, we'll take a closer look at exactly what some of the common wave shapes look like. 
First, though, we have to consider the relationship between n and I. Remember that a wave function for 
an electron always has three quantum numbers. In order to fully describe an electron wave, then, you have 
to know the values of all three of these numbers (n, I and mi). Now you might think that as long as n, 
I and nil are all integers, the wave that they describe will be a perfectly good electron wave. But that's 
NOT the way it works. It turns out that for a particular value of n, only certain values of I are allowed. 

Warning! For a particular value of n, only certain values of I are allowed. 

For wave functions that actually describe electrons in an atom, £ is always no less than 0, but also no more 
than n — 1(£ = 0,1,2, ...n— 1). The following examples should help to clarify the restriction on I. 

Example 1: 

What are the allowable values of I for an electron wave with n = 3? 
n = 3 

1 . Find the minimum value of I. 

The minimum value of t is always 0. 

2. Find the maximum value of €. 



The maximum value of I is always 1-1 
www.ckl2.org 



208 



maximum t = n — 1 
maximum t = 3 — 1 
maximum I = 2 

3. List a^ of the integers (no decimals!) starting from the minimum value of (, and ending with the 
maximum value of I. 

£ = 0,1,2 

Example 2: 

What are the allowable values of I for an electron wave with n = 1 
«= 1 

1 . Find the minimum value of (. 

The minimum value of I is always 0. 

2. Find the maximum value of t. 

The maximum value of I is always 1-1 
maximum i = 1 — 1 
maximum i = 

3. List a^ of the integers (no decimals!) starting from the minimum value of I and ending with the 
maximum value of t. 

£=0 

Often, scientists will refer the different values of t as electron sublevels. In a hydrogen atom with one 
electron, the value of t has no effect on the energy of the electron. However, in an atom with more than 
one electron, the value of I does have a small effect on the electron's energy. In other words, the principle 
quantum number, n, always determines the overall energy level of an electron, but that energy level is 
actually divided into multiple sublevels based on the value of t. All of these sublevels have equal energy 
in a hydrogen atom, because the hydrogen atom only has one electron. For atoms with more than one 
electron, though, the different sublevels split apart, and some of them end up having more energy than 
others. 

ml =Magnetic Quantum Number (Identifies Orbital) 

The third, and final quantum number, known as the magnetic quantum number, is given the symbol m\. 
Remember, the principal quantum number told you about the 'size' of an electron wave and the number 
of nodes in an electron wave, while the azimuthal quantum number told you more about the 'shape' of an 
electron wave. The magnetic quantum number, though, gives you even more information about what the 
electron wave looks like. The magnetic quantum number tells you how the electron wave is orientated in 
space. 

Orientation in space basically means where the electron wave points. Take a look at the two dumb-bell 
shaped electron waves shown in Figure 6.14 (in the next lesson, you'll learn that these are actually p 
orbitals). In the first electron wave, lobes of the wave point 'up-and-down' along the z-axis, while in the 

209 www.ckl2.org 





Figure 6.14: Two possible orientations for an electron wave with a dumb-bell shape. 

second electron wave, the lobes of the wave point 'in-and-out' along the x-axis. These two electron waves 
have different orientations in space, and thus different values for the quantum number m/. 

Now compare Figure 6.14 with Figure 6.15. In both figures, the electron wave has a different orientation 
in space, as indicated by the red arrow. But what do you notice about the orientations of the electron 
wave in Figure 6.15? They look the same, don't they? Obviously, orientation doesn't matter for the 
spherical electron wave. In other words, because the spherical electron wave looks the same no matter how 
you rotate it, there's really just one orientation. So how many different mi values should an electron wave 
with a spherical shape have (remember, mi values are used to describe orientation)? Clearly, an electron 
wave with a spherical shape should only have one mi value, because it only has one possible orientation. 
What about an orbital with a dumb-bell shape? Should it have a single mi value, or should it have several 
different m/ values? Well, since different orientations of the dumb-bell shaped wave actually look different, 
you'd probably expect the dumb-bell shaped wave to have several different m/ values, one for each possible 
orientation. In fact, the dumb-bell shaped wave actually has three possible values for mi, and so it has 
three possible orientations. (Don't worry about why there are exactly three orientations for the dumb-bell 
shaped wave. We'll talk a little bit about that in the next lesson, but the full explanation requires a lot of 
math that you'll learn about if you decide to study quantum physics or quantum chemistry). Hopefully, 
by comparing Figures 8 and 9, you should be convinced that the shape of the wave (which depends on () is 
important when it comes to determining the number of possible orientations (or the number of possible m/ 
values). It shouldn't be surprising, then, that for any given value of t, only certain mi values are allowed if 
you want to end up with a wave function that makes sense and actually describes electrons in a hydrogen 
atom. 

Warning! 

For a particular value of C, only certain values of m; are allowed. 

The rule for m/ is that for any value of (, mi can be any integer starting from — I and ending at +€. 
mi = -€, . . . , +i). The following examples should help to clarify the restriction on m/. 

Example 3: 

What are the allowable values of mi for an electron wave with t = 2? 



1. Find the minimum value of m/. 

The minimum, value of mi is always —I. 
www.ckl2.org 210 




y x 




Figure 6.15: Two possible orientations for an electron wave with a spherical shape. Notice that there isn't 
a noticeable difference between these orientations. 



minimum mi = -2 

2. Find the maximum value of m/. 

The maximum value of mi is always +£ 
maximum mi = +2 

3. List all of the integers (no decimals!) starting from the minimum value of mi and ending with the 
maximum value of m\ 

mi = -2,-1,0,1,2 

Example 4: 

What are the allowable values of m; for an electron wave with I = 0? 
£ = 

1 . Find the minimum value of mi 

The minimum value of mi is always —I. 
minimum m/ = 

2. Find the maximum value of m/. 

The maximum value of mi is always +t 
maximum mi = 

3. List all of the integers (no decimals!) starting from the minimum value of mi and ending with the 
maximum value of mi 

mi = 

211 www.ckl2.org 



Now that we've talked about all three of the different quantum numbers, you should have a good under- 
standing of how different electron waves can be described. You can describe the 'size' of an electron wave, 
and the number of nodes in an electron wave using the principal quantum number, n. You can describe 
the shape of an electron wave using the azimuthal quantum number, I. Finally, you can describe the 
orientation of the electron wave using the magnetic quantum number, m/. Since you know how to describe 
a general electron wave, it's time to look at several examples of specific electron waves. That's what we'll 
talk about in the next lesson. 

Lesson Summary 

• Schrodinger discovered that in order for a wave function to describe a standing wave that was con- 
tinuous, and that didn't 'doubled back' on itself, certain quantities in his wave function had to have 
integer values. 

The quantities in the wave function which must have integer values are known as quantum numbers. 
In the wave function for hydrogen, there are three quantum numbers. They are called the principal 
quantum number (n), the azimuthal quantum number (■£), and the magnetic quantum number (mi). 
The principal quantum number can only have positive integer values, (n = 1,2, 3 . . .). 
The principal quantum number determines how far the bulk of the electron density extends from the 
center of the atom. The higher the value of n, the further away from the nucleus you will be able to 
detect a significant amount of electron density. 

The principal quantum number also determines the number of nodes in an electron standing wave. 
The higher the value of n, the more nodes there are in the electron wave. 

The higher the principal quantum number, the greater the energy of the electron. Therefore the 
principal quantum number is determines the energy level of the electron. 

The azimuthal quantum number, (, determines the shape of the electron wave. The values of (, are 
also called the electron sublevels. They are labeled with the letters s, p, d, f, g, h, etc. 
For a wave function that actually describes an electron in an atom, t is always no less than zero, but 
also no more than n — 1(1 = 0, 1,2 . . . n— 1). 

In atoms with more than one electron, £ has a small effect on the electron's energy. 
The magnetic quantum number, m;, determines how the electron wave is orientated in space. For 
any given value of £, mi can be any integer from —I to + I (mi = —I, ■ ■■ , +€) . 

Review Questions 

1. Match each quantum number with the property that they describe. 

(a) n - \. shape 

(b) I - ii. orientation in space 

(c) mi - iii. number of nodes 

2. A point in an electron wave where there is zero electron density is called a . 

3. Choose the correct word in each of the following statements. 

(a) The (higher /lower) the value of n, the more nodes there are in the electron standing wave. 

(b) The (higher/lower) the value of n, the less energy the electron has. 

(c) The (more/less) energy the electron has, the more nodes there are in its electron standing wave. 

4. Fill in the blank. For lower values of n, the electron density is typically found the 

nucleus of the atom, while for higher values of n, the electron density is typically found - 

the nucleus of the atom. 

5. Circle all of the statements that make sense: Schrodinger discovered that certain quantities in the 
electron wave equation had to be integers, because when they weren't, the wave equation described 

www.ckl2.org 212 



waves which... 

(a) were discontinuous 

(b) were too small 

(c) were too long and narrow 

(d) were too short and fat 

(e) 'doubled back' on themselves 

6. What are the allowed values of t for an electron standing wave with n = 4? 

7. How many values of t are possible for an electron standing wave with n = 9? 

8. What are the allowed values of mi for an electron standing wave with t = 3? 

9. How many different orientations are possible for an electron standing wave with I = 4? 
10. What are the allowed values of mi for n = 21 

Further Reading / Supplemental Links 

• http : //www . Colorado . edu/physics/2000/elements_as_atoms/quantum_numbers . html 

• http : //dbhs . wvusd . kl2 . ca . us/webdocs/Electrons/QuantumNumbers . html 

• http : //www . wwnorton . com/college/chemistry/gilbert/concepts/chapter3/ch3_2 . htm 

• http : //www . sparknotes . com/testprep/books/sat2/chemistry/chapter4sect ion4 . rhtml 

Vocabulary 

quantum numbers Integer numbers assigned to certain quantities in the electron wave function. Be- 
cause electron standing waves must be continuous and must not 'double over' on themselves, quantum 
numbers are restricted to integer values. 

principal quantum number (n) Defines the energy level of the wave function for an electron, the size 
of the electron's standing wave, and the number of nodes in that wave. 

node A place where the electron wave has zero height. In other words, it is a place where there is no 
electron density. 

azimuthal quantum number {€) Defines the electron sublevel, and determines the shape of the electron 
wave. 

magnetic quantum number (m/) Determines the orientation of the electron standing wave in space. 

6.5 Shapes of Atomic Orbitals 

Learning Objectives 

• Define an electron orbital. 

• Be able to recognize s orbitals by their shape. 

• Be able to recognize p orbitals by their shape. 

213 www.ckl2.org 



Introduction 

In the last lesson, we learned how the principal quantum number determines the size of an electron wave 
(and the number of nodes), £ determines the shape of an electron wave, and m; determines the orientation 
of an electron wave. Now the effects of n are probably easy to visualize. For bigger values of n, the electron 
wave gets bigger, and it ends up with more nodes. Similarly, the effects of mi are easy to visualize as 
well. For different values of m/, the electron wave gets rotated into different orientations. In other words, 
the electron wave points in different directions. What about the effects of t ? You know that £ tells you 
something about the shape of the electron wave. You also know that certain waves are spherical while 
others are dumb-bell shaped, or butterfly shaped or just plain crazy shaped! But how do you know which 
value of £ corresponds to which shape? 

Unless you have a lot of training in mathematics, and can understand the wave function, there's really no 
way for you to predict what the shape of a wave with a particular value of £ will look like. Really, you just 
have to know, or be told - so that's what we'll do in the next few sections. We'll tell you what an £ = 
wave looks like, and what an £ = 1 wave looks like. 

When the Azimuthal Quantum Number is 0, The Electron Oc- 
cupies an s Orbital 

In the last lesson, you learned that different electron wave shapes have different names, and that these 
names are always letters of the alphabet like s,p,d and /. [These letters were chosen on the basis of 
observations of line spectra. Certain lines were observed as a "sharp" series; others were classified as " 
principal," "diffuse" or "fundamental series, thus s,p,d,f] These letters correspond to the shape of the 
electron wave, or at least the shape that the electron wave appears to take as a result of its electron density 
being thicker in one place than it is in the other. Now this can get a little confusing, but remember, electron 
waves describe electron density, or electron 'fog' which comes from interpreting the wave-like properties 
of the electron. Nevertheless, we always have to be careful that we don't forget about the particle-like 
properties of the electron as well. As a result, once scientists know the shape of a particular electron wave, 
they will often switch and start describing the electron as a particle again. To do this, they use the name 
for the shape of the wave, but rather than saying 'the electron wave has an s shape', or 'the electron 
wave has a p shape', they will say 'the electron is in an s orbital', or 'the electron is in a p orbital'. 
This makes it sound as if the electron is a particle again, and as if the orbital is some sort of box that the 
particle is confined to, or at least some sort of territory that the particle patrols. 

This switching back and forth between the wave-like description of the electron and the particle-like de- 
scription of the electron may seem confusing, or annoying, or just plain strange, but scientists do it to 
remind themselves that electrons are both particles and waves. So exactly what is an orbital? Technically 
speaking, an orbital is a wave function for an electron defined by the three quantum numbers, n, £ and 
mi. What the wave function describes, though, is a region in space with a particular shape, where you are 
likely to find an electron. In terms of waves, the orbital describes the region in space where the electron 
density is very thick. In terms of particles, the orbital describes the region in space where there is a high 
probability of finding the electron (which should make sense, because wherever the electron density is thick, 
there is also a high probability of finding the electron). 

So far we've decided that £ describes the shape of an orbital, which in turn describes where the electron 
density is thick (and there is a high probability of finding the electron) and where the electron density 
is thin (and there is a low probability of finding the electron). We're still, however, working towards a 
description of what these waves with regions of thick and thin electron density actually look like. Let's 
begin with the simplest wave shape. The simplest wave shape occurs for £ = 0. Whenever an electron wave 
is described by the quantum number £ = 0, we say that the wave function describes what's known as an s 

www.ckl2.org 214 



wave and thus the electron is in an s orbital, s waves can be big or small, depending on the value of n. s 
waves can also have a huge number of nodes, or no nodes at all, again depending on the value of n. What 
all s waves have in common, though, is their shape. That's what we'll talk about in the next section. 

s Sublevels are Spherically Shaped 

All I = electron waves are s waves, or waves from the s sublevel, and they all describe electrons in s 
orbitals. As suggested in the previous section, all electron waves from the s sublevel have the same overall 
shape, regardless of the value of n, regardless of their size, and regardless of the number of nodes they 
contain, s orbitals always correspond to spherical waves. The quantum numbers n = 1 and I = describe 
a small spherical wave with no nodes, the quantum numbers n = 2 and I = describe a larger spherical 
wave with a single node, and the quantum numbers n = 3 and I = describe an even larger spherical wave 
with two nodes. These waves all look slightly different, as shown in Figure 6.16. 






Figure 6.16: Various orbitals. All of these orbitals have , but they have different values for . The first 
orbital has , and thus is small and has no nodes. The second orbital has , and thus is larger and has one 
node. The third orbital has , and thus is even larger and has two nodes. 

Nevertheless, they are all spherical, because they all have I = 0. Their shapes don't change - only their 
sizes and the number of nodes that they contain. 

Now if you think back to an earlier lesson, you might remember something special about the different 
orientations of a spherical wave. Do you remember what happened when we rotated the spherical wave so 
that it pointed in different directions? It ended up looking the same, didn't it! No matter which way you 
rotate a sphere, it always looks the same. So how many different mi values do you expect for a spherical 
wave? One, of course! Now that you know spherical waves all have I = 0, you can use your rules for mi 
to figure out exactly how many different m/ are allowed. If you look back to Example 4 in the previous 
lesson, you'll see that we actually did that calculation. It turned out that there was only one allowable 
value for m/, and that was m/ = 0. In other words, there is only one orientation of a spherical wave. It all 
makes sense! 

So what does a spherical wave really mean? It means that your probability of finding an electron at any 
particular distance from the center of the atom only depends on the distance, and not on the direction. 
You can see this in Figure 6.17. 

Notice how it doesn't matter which direction you move as you travel from the center of the atom out, 
your probability of finding the electron is the same whether you head straight up, or straight down, or 
straight to the right, or straight to the left! The fact that electron density, and the probability of finding 
an electron is independent of direction is a special property of s orbitals, as you'll see shortly when we 
begin discussing some of the other possible electron orbitals like those for £ = 1. 



215 



www.ckl2.org 



1 L 



Figure 6.17: Notice that the amount of electron density (here represented by the intensity of the blue color) 
doesn't depend on direction. It does, however, depend on distance from the center of the atom. 

When the Azimuthal Quantum Number is 1, then ml Can Only 
Be -1, or +1 

All I = 1 electron waves are p waves, or waves from the p sublevel, and they describe electrons in what are 
known as p orbitals. Unlike s orbitals, p orbitals are not spherical, so they can have different orientations 
in space. Now that you know all p orbitals have t = 1, you should be able to figure out exactly how many 
different p orbital orientations exist by using your rules for m/. (m/ is the quantum number associated with 
the orientation of a particular orbital). Let's figure it out. 

Example 1: 

How many different p orbital orientations are possible? 

£= 1 

From now on, whenever you're told an electron is in a p orbital, you're expected to know that electron has 
the quantum number £ = 1 

The question how many p orbital orientations are possible, but what it's really asking is how many different 
mi values are allowed when ( = 1. We've already done this type of problem. 

1. Find the minimum value of mi. 

The minimum value of mi is always —I. 
minimum mi = -1 

2. Find the maximum value of mi. 

The maximum value of mi is always +t 
maximum mi = +1 

3. List all of the integers (no decimals!) starting from the minimum value of mi and ending with the 
maximum value of m/. 

mi = -1,0,1 

In this case m/ can equal -1,0 or 1, so there are a total of 3 allowed values for mi, and thus 3 possible p 
orbital orientations. 

www.ckl2.org 216 



The p Orbitals are Often Described as Dumb-bell Shaped 

Even though you know that there are three possible orientations for p orbitals, you can't really predict 
their shape unless you know a lot more about mathematics, physics and wave functions. When scientists 
use the wave function to draw the shape of an electron's p orbital, though, they always end up with is 
something that looks a lot like a dumb-bell. Not only that, the three different p orbitals (one with mi = — 1, 
another with mi = 0, and the third with mi = 1) turn out to be perpendicular to each other. In other 
words, if one p orbital points along the x-axis, another p orbital points along the y-axis, and the third 
points along the z-axis. Scientists typically label these three orbitals p x ,Py and p z respectively. Figure ?? 
shows each of the three p orbitals separately, and then all three together on the same atom. 









Sometimes we get so caught up thinking about electron wave functions, and electron orbitals, that we 
forget entirely about the atom itself. Remember, electron standing waves form because electrons get 
trapped inside an atom by the positive charge on the atom's nucleus. As a result, s orbitals, and p orbitals 
and even d and / orbitals always extend out from the atom's nucleus. Don't get so caught up in orbitals 
that you forget where they are and why they exist. 

As with s orbitals, p orbitals can be big or small, depending on the value of n, and they can also have 
more or less nodes, also depending on the value of n. Notice, however, that unlike the s orbital, which can 
have no nodes at all, p orbital always have at least one node. Take a look at the P orbital figure above 
again. Can you spot the node in each of the p orbitals? Since all p orbitals have at least one node, there 
are no p orbitals with n = 1. In fact, the first principal quantum number, n, for which p orbitals are 
allowed is n = 2. Of course you could have figured that out for yourself, right? No? Well, here's a hint - 
remember the rules for predicting which values of I are allowed for any given value of n. In the last lesson, 
you learned that t must be no less than 0, but also, no greater than n— 1. For the n = 1 energy level, then, 
the maximum allowed value for £ is 



maximum I = n -1 
maximum t = 1 — 1 
maximum I = 



217 



www.ckl2.org 



As a result, only s orbitals [i = 0) are allowed. For the n = 2 energy level, though, the maximum allowed 
value for I is 

maximum I = n — \ 
maximum t = 2 — 1 
maximum t — 1 

which means p orbitals {€ = 1) are allowed as well. So now you see that the restrictions on € are actually 
there to make sure that all n = 1 wave functions have no nodes, all n = 2 wave functions have 1 node, all 
n = 3 wave functions have 2 nodes, all... well, you get the picture. 

One interesting property of p orbitals that is different from s orbitals is that the total amount of electron 
density changes with both the distance from the center of the atom and the direction. Take a look at 
Figure ??. Notice how the electron density is different depending on which direction you travel from the 
center of the atom out. In the particular p orbital shown, the probability of finding the electron is greater 
as you head straight up from the center of the atom than it is as you head straight to the left or to the right 
of the atom. It turns out that this dependence on direction is very important when it comes to studying 
how different atoms interact and form bonds. We'll talk more about that in a later chapter. 

For p orbitals, the amount of electron density, and thus the probability of finding an electron, depends on 
both the distance from the center of the atom and the direction. (Source: CK-12 Foundation. CC-BY-SA) 

There is a high 
probability of finding the 
electron in this direction. 




There is very low < Tnere is ver V low 

probability of finding -^ |L ^- probability of finding 

the electron in this tne electron in this 

direction. direction. 



There is a high 
probability of finding the 
electron in this direction. 



The d Orbitals and f Orbitals are Not Easily Visualized 

Did you notice how the p orbitals looked a lot fancier than the simple spherical s orbitals? Well, you can 
imagine that if p orbitals with t = 1 are fancy, then d orbitals, with I = 2 are even fancier, and / orbitals, 
with I = 3 are just plain crazy! Most people can visualize p orbitals, but d orbitals and / orbitals are 
actually rather difficult to imagine. Most d-orbitals are butterfly shaped, although one has an unusual 
shape that looks like a donut surrounding a Q-tip! The 5 possible d orbitals are shown in Figure ?? 

www.ckl2.org 218 



(and you could have figured out that there were 5, right?). Don't worry too much about why one of the 
d orbitals is different. Again, it takes a lot of complex math to understand where the different d orbital 
pictures came from, and you won't have to worry about that unless you decide to go on and study quantum 
chemistry at the university level. As for / orbitals, they are even hard to draw, not to mention the fact 
that there are a total of 7 of them (Figure ??)! We won't be too concerned with d orbitals in this course, 
although they do become very important if you want to study certain metals like those found in the center 
of the periodic table. Similarly, / orbitals aren't all that important when it comes to common chemicals 
like hydrogen, or oxygen, or even copper. They do become important, though when you want to study 
some of the most important radioactive elements like uranium and plutonium! 

The probability patterns for the d orbitals. (Source: CK-12 Foundation. CC-BY-SA) 




x -y 



The probability patterns for the f orbitals. (Source: CK-12 Foundation. CC-BY-SA) 







V 



** 




rf A%^ 



219 



www.ckl2.org 



Information About the Quantum Theory 
General Information 

You should keep in mind that the theories scientists eventually accept are those that are in agreement with 
observations on nature and offer explanations for those observations. Theories about science that are the 
result of only mathematical development usually are not accepted until supportive observations are made. 
Such was certainly the case with Einstein's theory of relativity. 

In the case of quantum theory, even though it was developed primarily from observations on black-body 
radiation and light emissions from atoms, its application is much broader. The primary tenet of quantum 
theory is that energy comes in small packages called quanta. It doesn't say only light energy comes in 
packages, it says all energy comes in packages. One of these packets of energy is called a quantum (plural 
= quanta). In the specific case of light, a packet of energy can also be called a photon. Quantum theory 
says that all energy is quantized (comes in little bundles). That means that even the kinetic energy of 
a moving baseball must increase or decrease in steps. That is, a baseball may not have "any" possible 
energy but rather must have stepwise increases or decreases in energy. This stepwise change in energy is 
not possible to observe in baseballs, of course, because the change in the velocity of a baseball with the 
increase or decrease of one quantum of energy would be much too small for us to detect. In order for us 
to actually see the stepwise changes in energy of an object, the object must be extremely small so that a 
change in energy of a single quantum would be measureable. Of course, such observations are made in the 
gain or loss of energy by the electrons in the electron cloud of an atom when precise frequencies of light 
(corresponding to individual photons) are absorbed or emitted when electrons change energy levels. 

But, the stepwise change in energy is also detectable in the behavior of other very small objects. Molecules, 
in addition to their linear motion, also rotate. This rotational motion requires very small levels of energy 
and shows a stepwise increases and decreases in rotation. That is, certain molecules can be observed to 
rotate at 3 revolutions per second or 6 revolutions per second, or 9 revolutions per second, but cannot be 
observed to rotate at speeds in between those observed values. Molecules also have vibrational motion in 
their chemical bonds. This vibrational motion is also observed to be stepwise in nature. These are just 
some of the observations that are supportive of the quantum theory. 

Quantum Numbers and Electron Arrangement 

Quantum theory in general and specifically solutions to Schrodinger's equation gives us a great deal of 
information about the arrangement of electrons in the electron clouds of atoms. When we interpret this 
information, it is quite apparent that it came from a mathematical development. Keep in mind, however, 
that eventually you will see laboratory observations that have no explanation without quantum theory, 
just as the spectra of elements were observed 100 years before anyone could explain why it occurred. 

The principal quantum number, n, that comes from solutions to Schrodinger's equation indicates the major 
energy level that contains that electron. These energy levels are numbered 1,2,3,... etc. (The energy levels 
were originally named K,L,M,N,. . .etc, but later they were re-named 1,2,3,... and so on. Sometimes, in 
older text material, you will see references to the K energy level.) Since Schrodinger's equation is a 
theoretical statement, an infinite number of energy levels exist. However, when we place the electrons for 
our known atoms into the energy level structure, the electrons from even the largest known atom all fit in 
just 7 energy levels. There might be some theoretical discussion of an 8 or higher energy level but we 
never actually use any above 7. 

These energy levels also contain sub-levels. The sub-levels also have been named. The names of the sub- 
levels are s, p, d, and /. Mathematically, the number of sub- levels that an energy level can have is equal to 
the principal energy level number. That is, the 1 st energy level can have 1 sub-level, the 2 energy level 

www.ckl2.org 220 



can have 2 sub-levels, and so on. Obviously, if the number of sub-levels possible is equal to the energy level 
number, then energy level 6 could theoretically have 6 sub-energy levels. Therefore, some people list the 
sub-energy levels as s, p, d, f, g, h, i, . . . etc. Again, however, we find that our largest atom with the greatest 
number of electrons never uses any sub-level beyond f. Therefore, it is often said that the largest number 
of sub-levels used by atoms is 4. 

The sub-energy levels have the probability patterns for electrons that you saw in the last section. The 
s sub-energy level produces spherical probability patters, the p sub-levels produce the dumbbell shaped 
probability patterns, the d sub-energy levels produce the butterfly probability patterns, and the / sub-levels 
produce the probability patterns with no verbal description - but they do have a mathematical description. 
The n quantum number identifies the sub-level that contains the electron. When I = 0, the electron is in 
an s-orbital, when t = 1, the electron is in a p-orbital, when I = 2, the electron is in a J-orbital, and when 
I = 3, the electron is in an /-orbital. For all our known atoms, I never gets larger than 3. 

The sub-energy levels contain various spatial orientations of probability patterns called orbitals. This 
word is a carryover from the Bohr theory where an orbital was a circular path the electron followed around 
the nucleus but that is not the meaning of the word in modern chemistry. In modern chemistry, an orbital 
is a specifically designated volume inside the electron cloud where the probability of finding the electron 
is high. As mentioned earlier, the first energy level, n = 1, has only one sub-energy level because I must 
equal and that sub-level is an s sub-level. The number of orbitals in the s sub-levels was calculated in 
the last section. Sub-energy level s has a single orbital that is spherical in shape. Sub-energy level p has 
three dumbbell shaped orbitals, while sub-energy level d has five orbitals, and sub-energy level / has seven 
orbitals (Table 6.1). 

Table 6.1: Summary of Energy Levels, Sub- Levels, and Orbitals 

Energy Level Number Sub-Energy Levels Number of Orbitals 

1 s 1 

2 s 1 

2 p 3 

3 s 1 
3 p 3 

3 d 5 

4 s 1 
4 p 3 
4 d 5 

4 f 7 

5 s 1 
5 p 3 
5 d 5 
5 f 7 



For actual atoms, all the energy levels after 4 are repetitions of 4. That is, they have the same sub-energy 
levels and orbitals as energy level 4. In theory, of course, energy level 5 could have five sub-energy levels 
and the fifth sub-energy level would be named g and have 9 orbitals. We don't normally discuss that 
sub-energy level or any beyond it because they are never used. You might note the mathematical tone 
in this data, in that the total number of orbitals in any energy level is equal to the energy level number 
squared, n 2 . You probably also noticed that the number of orbitals in each succeeding sub-level increases 
by 2. 

221 www.ckl2.org 



Lesson Summary 

• An orbital is a wave function for an electron denned by the three quantum numbers, n, I and m/ . 
Orbitals define regions in space where you are likely to find electrons. 

• s orbitals {t = 0) are spherical shaped. 

• p orbitals (t = 1) are dumb-bell shaped. 

• The three possible p orbitals are always perpendicular to each other. 

Review Questions 

1. Fill in the blanks. When I = 0, the electron orbital is and when I = 1, the electron 

orbital is shaped. 

2. The n = 1 s orbital has nodes. 

3. The n = 2 s orbital has nodes. 

4. The n = 2 p orbital has nodes. 

5. The n = 1 p orbital has nodes. 

6. There are different p orbitals. 

7. What energy level (or value of n) has s, p and d orbitals, but no / orbitals? 

8. How many different d orbital orientations are there? 

9. How many / orbital orientations are there? 

10. How many different orbitals are there in the n = 3 energy level? 

Vocabulary 

orbital A wave function for an electron defined by all three quantum numbers, n, £, and mi. Orbitals 
define regions in space where there is a high probability of finding the electron. 



Image Sources 



(i 

(2 
(3 
(4 
(5 
(6 
(7 
(8 
(9 
(10 

(11 
(12 



Richard Parsons. . CC-BY-SA. 



Richard Parsons. . CC-BY-SA. 
Richard Parsons. . CC-BY-SA. 
http : //en . wikipedia . org/wiki/Image : Surf er_above_the_wave . jpg. GNU-FDL. 

Werner Heisenberg. Public Domain. 
CK-12 Foundation. . CC-BY-SA. 

http : //en . wikipedia . org/wiki/File : Conf ed-Cup_2005_-_Laolawelle . JPG. GNU-FDL. 



www.ckl2.org 222 



(13) A photograph of Erwin Schrodinger. Public Domain. 

(14) . 

(15) http://www.flickr.com/photos/davehogg/129247229/. CC-BY. 

(16) . 

(17) Louis de Broglie. Public Domain. 



223 www.ckl2.org 



Chapter 7 

Electron Configurations for 
Atoms 



7.1 The Electron Spin Quantum Number 

Lesson Objectives 

• Explain what is meant by the spin quantum number, m s . 

• Explain how the spin quantum number affects the number of electrons in an orbital. 

• Explain the difference between diamagnetic atoms and paramagnetic atoms. 

Introduction 

In the Quantum Mechanics Model of the Atom chapter, you learned about quantum numbers. Remember, 
quantum numbers actually come from the wave function. Even though you've never seen a mathematical 
equation for a wave function (and you'd probably think it was pretty scary if you did!) you can still 
understand what the different quantum numbers mean, because they all control different aspects of what 
the electron standing wave looks like. The principal quantum number, n, determines the 'size' and the 
number of nodes in the electron wave, the azimuthal quantum number, £, determines the 'shape' of the 
electron wave, and the magnetic quantum number, mi, determines the 'orientation' of the electron wave. 
Now suppose that you were told the exact size, the exact shape and the exact orientation of an electron 
orbital. Could you draw that object without being told anything else? You should be able to - what could 
you possibly need to know about an electron wave other than its size, shape and orientation? It seems 
as if the first three quantum numbers tell us all the information necessary to picture what an electron's 
probability pattern looks like - but it turns out that there's fourth quantum number! What! How could 
there be another quantum number? What property could it possibly describe? There's nothing left after 
size, shape, and orientation. That's exactly what scientists thought until they discovered spin. 

m s = +1/2 or -1/2 

In the early days, when scientists were just beginning to learn about quantum physics, most of what they 
knew about quantization involved atomic spectra. Remember, in The Bohr Model of the Atom chapter, you 
learned that atomic spectra were discontinuous, and that this led Bohr to propose the existence of quantized 
energy levels. To explain these quantized energy levels, Bohr suggested that electrons in an atom were 
restricted to specific orbits and that they moved around the nucleus in these orbits just like the planets 

www.ckl2.org 224 



move around the sun. It turned out that Bohr's model wasn't entirely correct because electrons don't 
orbit the nucleus. In fact, based on what scientists know of the wave-particle duality and the Heisenberg 
Uncertainty Principle, scientists now think that electrons can be anywhere inside the atom, although the 
probability of finding them at any one particular location over another depends on the relative amount 
of electron density at those locations. Even though Bohr's model had some problems, Bohr was right in 
his prediction that electrons could only exist at specific 'allowed' energy levels, and that all other energies 
were 'forbidden.' 

In the Quantum Mechanics Model of the Atom chapter, you learned how the 'allowed' and 'forbidden' 
energy levels could be explained in terms of electron standing waves. Only certain electron standing 
waves will fit perfectly inside the atom without 'doubling over' on themselves, or containing discontinuous 
jumps. Therefore, just like the Bohr Theory of the Atom, the wave function theory of the electron explains 
the existence of discontinuous atomic spectra. Actually, wave functions themselves are pretty good at 
predicting exactly what the atomic spectra of different atoms will look like (In other words, at which 
wavelengths lines of light will appear - if you have forgotten about atomic spectra, go back and read The 
Atomic Theory chapter!) Remember, Bohr's model could only predict the atomic spectrum of hydrogen. 
With a little bit of work, though, scientists could get electron wave functions to predict many different 
atomic spectra for many different types of atoms. It seemed perfect... but there was a still a slight problem. 
Every so often when scientists looked carefully at atomic spectra, they would find two separate lines of 
light at a wavelength where the wave function had only predicted one. The two separate lines were always 
really close together, so if you didn't look carefully, you'd think it was just one line and you'd think that 
the wave function was doing a great job. But the fact remained that frequently the wave function was 
missing some of the finer details of the atomic spectra. 

Now if you think back to what we discussed in The Bohr Model of the Atom chapter, you'll remember that 
the lines of light in an atomic spectrum result when an electron 'falls' from one energy level to another 
and releases its extra energy in the form of light. Even though we learned about this process in terms of 
the Bohr model, the same principle applies to the wave function theory as well. Remember, in terms of 
wave functions, the energy level of an electron is determined by the principal quantum number, n, while 
the azimuthal quantum number, {, defines the sublevel and also affects electron energy in atoms with more 
than one electron. 

Electrons don't circle the nucleus like planets in a solar system as pictured in the figure below on the left. 
Nevertheless, when they fall from a higher energy level electron standing wave to a lower energy electron 
standing wave they release light in much the same way as shown in the figure below on the right. 




When an electron begins at one energy (one value of n and I) and then falls to a lower energy (with a 
different value of n and £), it releases its extra energy in the form of light. Therefore, if there are two very 
closely spaced lines of light in an atomic spectrum, it must mean that there are two very closely spaced 
energy states from which (or to which) an electron can 'fall.' In terms of predicting the finer details of an 

225 www.ckl2.org 



atomic spectrum, then, it's clear that the problem with the wave function was that in certain situations it 
would only find one allowed energy state where there should have been two! 




Figure 7.1: George Uhlenbeck (on the far left) and Samuel Goudsmit (on the far right), the two scientists 
who first proposed the existence of the spin quantum number. 



The story got even stranger when scientists put magnets around different atoms and looked at what 
happened to their atomic spectra as a result. All of those pairs of closely spaced spectral lines that the 
wave function couldn't predict actually split apart even further! In other words, magnetic fields affected 
the two different energy levels differently. It made the low energy level drop even lower, and it made the 
high energy level rise even higher. After the magnet experiments, scientists came to the realization that 
their wave function description was not as complete as they had hoped. There was some other property 
that had to be included - but what? 

In 1925, two scientists by the names of Samuel Goudsmit and George Uhlenbeck (Figure 7.1) suggested 
extending the wave function equation so that it included a fourth quantum number, called the 'spin' 
quantum number. When any charged object (like, for example, a negatively charged electron) spins in 
a magnetic field, its energy is determined by the direction in which it rotates. In other words, if the 
object rotates clockwise, it will have a different amount of energy than it would have if it had rotated 
counterclockwise. Goudsmit and Uhlenbeck argued that if the electron was spinning, then it was easy to 
explain why those two closely spaced spectral lines split further apart in the presence of a magnetic field. 

One line corresponded to the energy level for the electron spinning clockwise, and the other corresponded 
to the energy level for the electron spinning counterclockwise. (If you're really clever, you might wonder 
why these two energy levels were different and could be seen in atomic spectra even when there weren't 
any external magnets around. It turns out that there are always tiny internal magnetic fields created by 
matter itself. Usually, though, these internal magnetic fields are very small, so their effect is very small as 
well.) 

Pictured below are two negatively charged particles spinning in a magnetic field. The left particle will 
have a different energy than the right particle, because they are spinning in different directions. (Source: 
CK-12 Foundation. CC-BY-SA) 

www.ckl2.org 226 



negatively charged particle 
spinning counterclockwise 




negatively charged particle 
spinning clockwise 



magnetic field 
* V V t V 

■ 



By incorporating electron spin into the electron wave function, scientists found that the fourth quantum 
number, also known as the spin quantum number, m s , could take on two different values - they were 
m s = +1/2 and m s = -1/2. Notice that unlike n, I and mi, which can only be integers, m s can be half- 
integers. The value of m s will never be some crazy decimal like m s = 0.943895, but it can be the decimal 
m s = 0.5, because that corresponds to one-half, or a half-integer. 

Now it's tempting to think of m s = +1/2 as the electron spinning in one direction, and m s = -1/2 as the 
electron spinning in the other direction. When you do that, though, how are you picturing the electron? 
You're picturing it as a particle, aren't you? And how do you explain the wave-like nature of an electron 
in terms of rotation clockwise or counterclockwise? Well, you can't. So despite the fact that the idea for 
including electron spin in the wave function came from picturing electrons as tiny little spinning objects, 
scientists try not to make any direct comparisons between the spin quantum number and a particle rotating 
clockwise or counterclockwise. Instead, scientists usually make vague statements like 'spin can only truly 
be understood as a quantum property. There is no directly comparable macroscopic (large scale) property'. 
Rather than saying anything about a clockwise rotation, or a counterclockwise rotation, scientists call 
electrons with m s = +1/2 'spin-up' and electrons with m s = -1/2 'spin down'. 



When Two Electrons Occupy the Same Orbital, They Will Have 
Opposite Spins 

Do you remember what an orbital is? An orbital is a wave function for an electron with a specific set of 
quantum numbers, n, I, and ra/. For example, the numbers n = 1, £ = and m/ = define one particular 
orbital, while the numbers n = 2, £ = 1 and mi = -1 define another orbital, and the numbers n = 2, I = 1 
and mi = define a different orbital again. Orbitals are very important because any time you know the 
values of the first three quantum numbers n, £ and mi, you know the region in space where there is a high 
probability of finding the electron. In other words, if you know which orbital an electron is in, you know 
exactly what its 'box', or 'territory' looks. The orbital is really just a description of where the electron 
spends most of its time. 



227 



www.ckl2.org 



For example, if you're told that the electron is in an orbital with n = 1, I = 0, mi = 0, you automatically 
know that the electron is probably going to be found in a spherical shaped 'territory' that's pretty close 
to the nucleus of the atom. On the other hand, if you're told that the electron is in an orbital with n = 5, 
I = 1 and mi = you know that the electron is probably going to be found in a dumb-bell shaped 'territory' 
that extends quite far from the nucleus of the atom, along the y axis (technically speaking mi = could 
be the orbital along the x-axis, the orbital along the y-axis or the orbital along the z-axis, but whichever 
axis you choose for the m; = orbital, the mi = 1 and mi = -1 orbitals will automatically point along the 
other two axes). 

The spin quantum number doesn't tell you anything about the region in space where you're likely to find 
the electron. Therefore, if you want to know which orbital an electron is in, all you need to know is the 
values of the first three quantum numbers - you don't need to know the value of the fourth. So why is the 
value of the fourth quantum number important? Well, the spin quantum number determines whether or 
not an electron is allowed to share an orbital with another electron that's already there. It turns out that 
each region of the atom (or each orbital) can actually be shared by two electrons. 

Two electrons can share their territory, or their orbital, provided each electron does something different. 
This means that sharing an orbital is only possible for electrons which have different spin quantum numbers. 
An electron that has a spin quantum number of m s = +1/2 can share an orbital with an electron that has a 
spin quantum number of m s = -1/2. However, an electron that has a spin quantum number of m s = +1/2 
cannot share an orbital with another electron that also has a spin quantum number of m s = +1/2. Similarly, 
an electron that has a spin quantum number of m s = -1/2 cannot share an orbital with another electron 
that has a spin quantum number of m s = -1/2. 

When Electrons are Paired, They are Diamagnetic i.e. No Mag- 
netic Attraction 

In the last section, you learned that any time two electrons share the same orbital, their spin quantum 
numbers have to be different. In other words, one of the electrons has to be 'spin-up', with m s = +1/2, 
while the other electron is 'spin-down', with m s = -1/2. This is important when it comes to determining 
the total spin in an electron orbital. Technically speaking, an electron's spin isn't exactly the same as its 
spin quantum number, but the difference isn't important when it comes to figuring out whether or not two 
electron spins cancel out. In order to decide whether or not electron spins cancel all you need to do is 
add their spin quantum numbers together. If the total is 0, then the spins cancel each other out. If the 
total is greater than 0, or less than 0, then the spins do not cancel each other out. Let's take a 
look at two examples: 

Example 1: 

Electron A has m s = +1/2 while electron B has m s = -1/2. Do the spins of these two electrons cancel each 
other out? 

m s for A = +1/2 

m s for B = -1/2 

Total m s = m s for A + m s for B 
Total m s = (+1/2) + (-1/2) 
Total m s = 1/2-1/2 
Total m s = 

In this case, the spins of electron A and electron B do cancel each other out. 
www.ckl2.org 228 



Example 


2: 


Electron A has m s 


other out? 




m s for A = 


-1/2 


m s for B = 


-1/2 



-1/2 while electron B has m s = -1/2. Do the spins of these two electrons cancel each 



Total m s = m s for A + m s for B 
Total m s = (-1/2) + (-1/2) 
Total m s = -1/2 - 1/2 
Total m. = -1 



In this case, the spins of electron A and electron B do not cancel each other out. 

The idea of 'canceling out' should make sense to you. If one spin is 'up,' and the other is 'down,' then the 
'up' spin cancels the 'down' spin and there is no leftover spin at all. The same logic applies if we think 
of spins as 'clockwise' and 'counterclockwise.' If one spin is 'clockwise' and the other is 'counterclockwise,' 
then the two spin directions balance each other out and there is no leftover rotation at all. Notice what 
all of this means in terms of electrons sharing an orbital. Since electrons in the same orbital always have 
opposite values for their spin quantum numbers, m s , they will always end up canceling each other out! In 
other words, there is no leftover (or 'net') spin in an orbital that contains two electrons. Whenever two 
electrons are paired together in an orbital, we call them diamagnetic electrons. On the other hand, an 
orbital containing only one electron will have a total spin equal to the spin of the electron that it contains. 

Even though electron spin can only be truly understood using quantum physics, it does produces effects 
that we can actually see in our everyday lives. Electron spin is very important in determining the magnetic 
properties of an atom. If all of the electrons in an atom are paired up and share their orbital with another 
electron, then the total spin in each orbital is zero and, by extension, the total spin in the entire atom is zero 
as well! When this happens, we say that the atom is diamagnetic because it contains only diamagnetic 
electrons. Diamagnetic atoms are NOT attracted to a magnetic field. In fact, diamagnetic atoms are 
slightly repelled by magnetic fields. 

Since diamagnetic atoms are slightly repelled by a magnetic field, it is actually possible to make certain 
diamagnetic materials float! Below you can see a thin black sheet of pyrolytic graphite floating above the 
gold magnets. (Source: http://en.wikipedia.Org/wiki/Image:Diamagnetic_graphite_levitation. 
jpg. Public Domain) 

229 www.ckl2.org 




When an Electron is Unpaired, Paramagnetism will be Observed 



If atoms that contain only paired electrons are slightly repelled by a magnetic field, what do you think 
is true of atoms that contain unpaired electrons? If you guessed that atoms with unpaired electrons are 
slightly attracted to a magnetic field, then you guessed right! Whenever electrons are alone in an orbital, 
we call them paramagnetic electrons. Remember, if an electron is all by itself in an orbital, the orbital 
has a 'net' spin, because the spin of the lone electron doesn't get canceled out. If even one orbital has a 
'net' spin, the entire atom will have a 'net' spin as well. Therefore, we say an atom is paramagnetic 
when it contains at least one paramagnetic electron. Notice that the definition of diamagnetism and 
the definition of paramagnetism are subtly different. This can be confusing if you aren't careful. Be sure 
to note: In order for an atom to be diamagnetic, all of its electrons must be paired up in orbitals. In order 
for an atom to be paramagnetic, at least one of its electrons must be unpaired. 

In other words, an atom could have 10 paired (diamagnetic) electrons, but as long as it also has 1 unpaired 
(paramagnetic) electron, it's still considered a 'paramagnetic atom'. In order to be a 'diamagnetic atom', 
the atom would have to have 10 paired (diamagnetic) electrons and no unpaired (paramagnetic) electrons. 
Just as diamagnetic atoms are slightly repelled from a magnetic field, paramagnetic atoms are 
slightly attracted to a magnetic field. 

Lesson Summary 

• If you only consider the first three quantum numbers, the wave function model for the electron will 
sometimes predict one spectral line where there are actually two closely spaced spectral lines. 

• This led to the proposal of a fourth quantum number, the spin quantum number m s . 

• fflj can have two possible values for an electron. It can be 'spin-up' with m s = +1/2 or 'spin-down' 
with m s = -1/2 

• When two electrons occupy the same orbital, they must have different spin quantum numbers. 
www.ckl2.org 230 



• An orbital containing two electrons will have no net spin. When this is the case, the two electrons 
are called diamagnetic electrons. 

• An orbital containing only one electron will have a total spin equal to the spin of the electron that 
it contains. When this is the case, the electron is called a paramagnetic electron. 

• Electron spin helps to determine the magnetic properties of an atom. 

• If all electrons in an atom are diamagnetic, the entire atom has no net spin, and is termed a 'dia- 
magnetic atom.' Diamagnetic atoms are slightly repelled from a magnetic field. 

• If an atom contains even one paramagnetic electron, the entire atom has a net spin and is termed a 
paramagnetic atom. Paramagnetic atoms are slightly attracted to a magnetic field. 

Review Questions 

1. The principal quantum number describes the size of an electron energy level, the azimuthal quantum 
number describes the shape of an electron energy level, and the magnetic quantum number describes 
the orientation of the electron energy level. If there was another quantum number, what do you 
think it might describe about the electron? 

2. There is, in fact, a fourth quantum number that we'll learn about in this lesson. The fourth quantum 
number is called the spin quantum number. Now can you guess what the final quantum number 
might describe? 

3. Choose the correct statement. 

(a) The spin quantum number for an electron can only have the values m s = +1 and m s = — 1 

(b) The spin quantum number for an electron can only have the value m s = 

(c) The spin quantum number for an electron can have any integer value between —I and +( 

(d) The spin quantum number for an electron can only have the values m s = +1/2 and m s = -1/2 

(e) The spin quantum number does not apply to electrons 

4. Choose the correct statement. 

(a) When two electrons share an orbital, they always have the same spin quantum numbers 

(b) When two electrons share an orbital, they always have opposite spin quantum numbers 

(c) Two electrons cannot share the same orbital 

(d) When two electrons share an orbital there is no way to predict whether or not they will have 
the same spin quantum numbers 

5. Fill in the blanks in the following statement using numbers. 

When scientists used the Schrodinger equation with only quantum numbers, they found that the 

Schrodinger equation was pretty good at predicting atomic spectra, except that there were occasionally 

closely spaced lines of light where the Schrodinger equation predicted only . This led scientists to 

suggest that a complete description of an electron, which required quantum numbers. 

6. In many atomic spectra, there are two very closely spaced lines of light which can only be predicted 
by including the spin quantum number into the Schrodinger equation. Decide whether the following 
statements about these two lines are true or false. 

(a) the two lines spread further apart when the atom is placed in a magnetic field 

(b) the two lines move closer together when the atom is placed in a magnetic field 

231 www.ckl2.org 



(c) the two lines are the result of an experimental error. If scientists are careful, they find that 
there is really just one line. 

(d) the two lines actually result from the fact that there are two very closely spaced energy states 

7. Goudsmit and Uhlenbeck proposed the existence of 

(a) the principal quantum number 

(b) the azimuthal quantum number 

(c) the spin quantum number 

(d) the magnetic quantum number 

8. Circle all of the quantum numbers that tell you about the region in space where you're most likely 
to find the electron. 

(a) the spin quantum number 

(b) the magnetic quantum number 

(c) the principal quantum number 

(d) the azimuthal quantum number 

9. Select the correct statement from the list below. An electron with a spin quantum number of 
m s = -1/2 

(a) cannot share an orbital with an electron that has a spin quantum number of m s = +1/2 

(b) prefers to share an orbital with an electron that has a spin quantum number of m s = -1/2 

(c) cannot share an orbital with an electron that has a spin quantum number of m s = -1/2 

(d) cannot share an orbital with another electron 

10. What is the total spin in an electron orbital if 

(a) the orbital contains one 'spin-up' electron 

(b) the orbital contains one 'spin-down' electron 

(c) the orbital contains two 'spin-up' the orbital contains one 'spin-up' electron and one 'spin-down' 
electron 

(d) . 

Further Reading / Supplemental Links 

• http : //www . ethbib . ethz . ch/exhibit/pauli/elektronenspin_e . html 

• http : //www . lorentz . leidenuniv . nl/history/spin/goudsmit . html 

• http://en.wikipedia.org/wiki 

Vocabulary 

spin quantum number, m s The fourth quantum number that must be included in the wave function of 
an electron in an atom in order to completely describe the electron. 

spin-up The term applied to electrons with spin quantum number m s = +1/2. 

spin down The term applied to electrons with spin quantum number m s = -1/2. 

diamagnetic electrons Two electrons with opposite spins, paired together in an orbital. 
www.ckl2.org 232 



diamagnetic atom An atom with no net spin; an atom with only diamagnetic electrons. 

paramagnetic electron An electron alone in an orbital. 

paramagnetic atom An atom with a net spin; an atom with at least one paramagnetic electron. 

7.2 Pauli Exclusion Principle 

Lesson Objectives 

• Explain the meaning of the Pauli Exclusion Principle. 

• Determine whether or not two electrons can coexist in the same atom based on their quantum 
numbers. 

• State the maximum number of electrons that can be found in any orbital. 

Introduction 

When electrons are found inside an atom, they're restricted to specific areas, or regions within the atom 
which can be described by orbitals. Let's see what this means in terms of quantum numbers. 

No Two Electrons in an Atom Can Have the Same Four Quantum 
Numbers 

How do you know that two electrons are in the same orbital? In order to fully specify an orbital, you need 
to know the principal quantum number, n, the azimuthal quantum number, C, and the magnetic quantum 
number, m\. The values of first three quantum numbers for an electron determine exactly which orbital the 
electron in. Clearly, then, in order to be in the same orbital, two electrons have to have exactly the same 
values for n, {, and m/. Now when two electrons have exactly the same values for n, t, and mi, they share 
the same region of space within the atom, and in the last lesson, you learned that that had important 
consequences in terms of their spins. If you remember back to an earlier section, electrons in the same 
orbital, sharing the same region of space, had to have different values of m s . If one electron had m s = +1/2, 
then the other had to have m s = -1/2 and vice versa. Let's take a look at several examples. 

Example 1: 

An electron with n = 2, t = 1, mi = — 1 and m s = +1/2 is found in the same atom as a second electron 
with n = 2, t = 1, and m; = — 1. What is the spin quantum number for the second electron? 

First electron: n = 1,1 = l,m\ = — l,m s = +1/2 

Second electron: n = 1,1 = l,m/= —l,m s =? 

Since the first three quantum numbers are identical for these two electrons, we know that they are in the 
same orbital. As a result, the spin quantum number for the second electron cannot be the same as the spin 
quantum number for the first electron. This means that the spin quantum number for the second electron 
must be m s = -1/2. 

Example 2: 

An electron with n = 5, I = 4, m/ = 3 and m s = -1/2 is found in the same atom as a second electron with 
n = 5, t = 4, and m/ = 3. What is the spin quantum number for the second electron? 

233 www.ckl2.org 



First electron: n = 5,-C = 4, mi = 3,m s = -1/2 

Second electron: n = 5, t = 4, m; = 3,m s =? 

Since the first three quantum numbers are identical for these two electrons, we know that they are in the 
same orbital. As a result, the spin quantum number for the second electron cannot be the same as the spin 
quantum number for the first electron. This means that the spin quantum number for the second electron 
must be m s = +1/2. 

Notice that whenever the two electron's first three quantum numbers are the same, the fourth is different. 
Let's take a look at a few more examples... 

Example 3: 

Can an electron with n = 1, t = 0, mi = and m s = +1/2 exist in the same atom as a second electron with 
n = 2, I = 0, mi = and m, = +1/2? 

First electron: n = 1,1 = 0,w/ = 0,m s = +1/2 

Second electron: n = 2,1 = 0, mi = 0,m s = +1/2 

Since these two electrons are in different orbitals, they occupy different regions of space within the atom. 
As a result, their spin quantum numbers can be the same, and thus these two electrons can exist in the 
same atom. 

Example 4: 

Can an electron with n = 3, I = 1, mi = — 1 and m s = -1/2 exist in the same atom as a second electron 
with n = 3, t = 2, mi = — 1 and m s = -1/2? 

First electron: n = 3,£ = l,mi = —l,m s = -1/2 

Second electron: n = 3, € = 2, mi = -l,m s = -1/2 

Since these two electrons are in different orbitals, they occupy different regions of space within the atom. 
As a result, their spin quantum numbers can be the same, and thus these two electrons can exist in the 
same atom. 

Example 5: 

Can an electron with n = 1, £ = 0, mi = and m s = +1/2 exist in the same atom as a second electron with 
n = 2, t = 1, mi = and m s = +1/2? 

First electron: n = 1,1 = 0,ra/ = 0,m s — +1/2 

Second electron: n = 2, 1 = \,m\ = 0,m s = +1/2 

Since these two electrons are in different orbitals, they occupy different regions of space within the atom. 
As a result, their spin quantum numbers can be the same, and thus these two electrons can exist in the 
same atom. 

Notice that whenever the two electrons have different values of n, or different values of {, or different values 
of mi, they can have the same spin quantum number m s , because they are not in the same orbital, and 
thus they are not sharing the same region of space within the atom. Let's take a look at one final example 

Example 6: 

Can an electron with n = 1, t = 0, mi = and m s = +1/2 exist in the same atom as a second electron with 
n = l ; l = 0, mi = and m, = +1/2? 

First electron: n = 1,1 = 0,ra/ = 0,m s = +1/2 

Second electron: n = 2, 1 = l,m; = 0,m s = +1/2 

Since these two electrons are in the same orbital, they occupy the same region of space within the atom. 

www.ckl2.org 234 



As a result, their spin quantum numbers cannot be the same, and thus these two electrons cannot exist in 
the same atom. 

Hopefully after having looked at seven different examples, it should be obvious to you that electrons in the 
same atom with the same spin must be in different orbitals, while electrons in the same orbital of the same 
atom must have different spins. As a result, no two electrons in the same atom can have exactly the same 
four quantum numbers. If two electrons have the same n, the same €, and the same mi, then they are in 
the same orbital. If they also have the same m s , then they also have the same spin, and that is impossible. 

The first scientist to realize that two electrons in the same atom couldn't have the same four quantum 
numbers was a man name Wolfgang Pauli (Figure 7.2). In 1925, Pauli stated what has come to be known 
as the Pauli Exclusion Principle. The Pauli Exclusion Principle states that no two identical fermions 
(a fancy word for electrons and other subatomic particles like electrons) may occupy the same quantum 
state in an atom simultaneously. In other words, no two electrons in the same atom can have the same 
four quantum numbers. If n, I, and mi are the same, m s must be different such that the electrons have 
opposite spins. 





* /is 


■% 
t 

s 

0- 













r 





Figure 7.2: Wolfgang Pauli, the scientist who first proposed the Pauli Exclusion Principle. 



No Atomic Orbital Can Contain More than Two Electrons 

An electron can share its territory, or its orbital, with another electron, but only if the other electron is 
slightly different - in other words, only if the other electron has a different spin. 

There's a limit to the number of different electrons that can share an orbital, because there's a limit to the 
number of different spins that those electrons can have. When it comes to spins, though, there are only 
two possibilities. An electron can either be 'spin-up', with m s = +1/2, or 'spin-down', with m s = -1/2. 
Therefore, if an orbital has one electron that is 'spin- up', and a second electron that is 'spin-down', the 
orbital is full. What if a third electron tried to enter the orbital? Well, if the third electron was 'spin-up' 
it would have trouble sharing the orbital, with the 'spin- up' electron that's already there. Similarly, if 
the third electron was 'spin-down', it would have trouble sharing the orbital with the 'spin-down' electron 
that's already there. Since the only two options for the third electron are 'spin-up' and 'spin-down', there's 
really nothing that third electron can do - it just has to move on and find a new orbital! To summarize, 
then, because there are only two possibilities for the spin quantum number of an electron, 

NO ATOMIC ORBITAL CAN CONTAIN MORE THAN TWO ELECTRONS! 

235 www.ckl2.org 



Lesson Summary 



• The Pauli Exclusion Principle states that 'no two identical fermions may occupy the same quantum 
state in an atom simultaneously'. That is, no two electrons in an atom can have n, t, mi and m s all 
the same. 

• No atomic orbital can contain more than two electrons. 

Review Questions 

1. Electrons in the same orbital must have different spin quantum numbers. What is true of the other 
three quantum numbers for two electrons in the same orbital? 

2. Electrons in different orbitals can have the same spin quantum numbers. What is true of the other 
three quantum numbers for two electrons in different orbitals? 

3. Fill in the blank using either the word 'can', or 'cannot'. 

(a) An electron with the quantum numbers n = 1, I = 0, m/ = and m s = +1/2 exist in 

the same atom as an electron with the quantum numbers n = 2, £ = 0, m; = and m s = +1/2 

(b) An electron with the quantum numbers n = 1, € = 0, mi = and m s = +1/2 exist in 

the same atom as an electron with the quantum numbers n = 1, I = 0, mi = and m s = -1/2 

4. Fill in the blanks using numbers. 

(a) There is only 1 orbital at the n = 1 energy level. Therefore the n = 1 energy level can hold a 
maximum of electrons 

(b) There are 4 orbitals at the n = 2 energy level. Therefore the n = 2 energy level can hold a 
maximum of electrons 

(c) There are 9 orbitals at the n = 3 energy level. Therefore the n = 3 energy level can hold a 
maximum of electrons 

(d) There are 16 orbitals at the n = 4 energy level. Therefore the n = 4 energy level can hold a 
maximum of electrons 

5. What is the maximum number of electrons that can exist in p orbitals at energy levels with n < 3. 

6. What is the maximum number of electrons that can exist in p orbitals at energy levels with n < 5. 

Further Reading / Supplemental Links 

• http : //theory . uwinnipeg . ca/mod_tech/nodel68 . html 

• http://en.wikipedia.org/wiki 

Vocabulary 

Pauli Exclusion Principle No two fermions may occupy the same quantum state in an atom simulta- 
neously; no two electrons in an atom can have the same four quantum numbers. 

7.3 Aufbau Principle 

Lesson Objectives 

• Explain the Aufbau Principle. 

• Given two different orbitals, predict which the electron will choose to go into. 

www.ckl2.org 236 



Introduction 

Everything in the universe is driven to minimize its potential energy. Previously we learned that heavy 
objects fall when they're dropped, because their total potential energy is lower on the ground than it is 
when they're hovering in the air. That's why Wile E. Coyote could drop an anvil on Road Runner - 
he knew that the anvil would fall, because falling would lower its potential energy. It's the same with a 
bowling ball placed on a hill. At the top of the hill, the bowling ball has more potential energy than it 
does at the bottom, so you can always bet on the bowling ball rolling down the hill rather than up the hill, 
because the bowling ball will always try to minimize its energy. In this chapter, we won't look at heavy 
objects like bowling balls and anvils. Instead, we'll look at tiny objects - we'll look at electrons. Even 
though electrons are much, much smaller than anvils and bowling balls, the same principle applies. An 
electron will do anything that it can to lower its potential energy. 

Electrons are Found in Energy Levels with Increasingly Higher 
Energy 

Whenever an electron is found inside an atom, it exists in what's known as an orbital. By now, you 
should know what an orbital is. An orbital describes a particular region of space within the atom where 
the electron is most likely to be found. While orbitals are important when it comes to figuring out an 
electron's probable location, they are equally important when it comes to figuring out an electron's energy. 
Electrons in different orbitals frequently have different energies. Of course, an electron is never going to 
"choose" to be in an orbital that has a higher energy if there's space available in an orbital that has a 
lower energy. It's a lot like you riding the bus. You would never choose to waste energy walking 20 miles 
to school provided there was space for you on the bus. If there wasn't any space on the bus, though, you 
may be forced to walk. It's the same with electrons. 

As pictured below, if there's no room on the bus, you may be forced to walk. It's the same with electrons. 
If there's no room in a low energy orbital, they may be forced into a higher energy orbital. (Source: 
http : //en . wikipedia . org/wiki/Image : HDaumierOmnibus . JPG . Public Domain) 




If there isn't any space in a low energy orbital, an electron may be forced into a higher energy orbital. The 
fact that electrons always fill up lower energy orbitals first has important consequences when it comes to 
determining which orbitals contain electrons in any given atom. 



237 



www.ckl2.org 



Remember that the principal quantum number, n, is associated with the 'energy level' of the electron. 
Electrons with standing waves described by bigger values of n had higher energies, while electrons with 
standing waves described by smaller values of n had lower energies. If you were an electron, then, and 
you had the choice of being in an orbital with n = 1 or an orbital with n = 2, which would you choose? 
Obviously you'd choose to be in the orbital with n = 1, because it has a smaller value of n, and thus a 
lower energy (remember, both people and electrons prefer to be in states with lower energy). 

Naturally, there is a limit to the total number of electrons that can exist in the same atom at the n = 1 
energy level. In fact, it turns out that there can be at most two electrons with n = 1 in any given atom. 
That's because there is only one n = 1 orbital per atom. Of course, there are many atoms with more than 
two electrons. Lithium, for instance, has three. What happens to the electrons in an atom like lithium? 
Obviously, the first two electrons are going to occupy the single orbital that exists at the n = 1 energy 
level. Since this orbital only has room for two electrons, though, the third electron has to move up to the 
n = 2 energy level. In other words, electrons will fill up orbitals in order of increasing energy. If there's 
space at the n = 1 energy level, that space will be filled before any electrons move into the n = 2 energy 
level. Similarly, if there's space at the n = 2 energy level, that space will be filled before any electrons 
move into the n = 3 energy level. 



In an Energy Level Electrons are Assigned to Sublevels with In- 
creasingly Higher Energy 

So far you know that an electron will always be found occupying the orbital in the first energy level, rather 
than the second energy level, provided that there is space available. Similarly, an electron will be found 
occupying the second energy level, rather than the third energy level, provided that there is space available. 
But what about electron sublevels? Among the orbitals with the same value of n, the s orbital will always 
be filled first, followed by the p orbitals (and then the d orbitals). 

We can summarize these rules with the following statement: Electrons will fill available orbitals starting 
with those at the lowest energies before moving to those at higher energies. This statement is known as 
the Aufbau Principle. That may sound like a funny name, but 'aufbau' is actually the German word 
for 'construction', and the Aufbau Principle describes how the orbitals are 'constructed' by progressively 
adding electrons to higher and higher energy levels. 

Before we move on and consider exactly how many electrons go into each energy level, and sublevel, it's 
important to point out how the energies of the electrons in an atom relate to the energy of the entire atom 
itself. If each electron tries to minimize its energy by going into the lowest energy orbital available, then 
the total energy of all the electrons in the atom is also as low as possible. 

Let's compare electrons minimizing their energy in an atom to your relatives minimizing their energy 
consumption in your family. If you do your best to turn off the lights and save as much energy as possible, 
and your brother does his best to turn off the lights and save as much energy as possible, and your sister 
does her best to turn off the lights and save as much energy as possible, and your parents also do their best 
to turn off the lights and save as much energy as possible, then collectively, your entire family is saving as 
much energy as possible as well. It's the same with electrons. If the first electron fills the lowest energy 
orbital available to it, and the second electron fills the lowest energy orbital available to it, and the third 
electron fills the lowest energy orbital available to it, and the fourth electron also fills the lowest energy 
orbital available to it, then collectively, the entire atom is in the lowest energy state possible as well. 

Just as you would prefer to minimize your energy by sleeping, or riding a bus, electrons minimize their 
energy by occupying the lowest energy orbital available to them and atoms minimize their energy by having 
all of their electrons in the lowest energy 'configuration' (arrangement) possible. Later, you will learn that 
practically all chemical processes rely on this same principle of energy minimization. 



www.ckl2.org 238 



Lesson Summary 

• In an atom, electrons will fill up orbitals in order of increasing energy. 

• The principle quantum number determines the 'energy level' of the orbital. Orbitals with lower values 
of n are usually associated with lower energy and will be filled first. 

• The azimuthal quantum number determines the 'sublevel' of the orbital. 

• Orbitals with lower values of t (but the same value of n) are always associated with lower energy and 
will be filled first. 

• The Aufbau Principle states that electrons will fill available orbitals starting with those at the lowest 
energies before moving to those at higher energies. 

• Since each electron in an atom minimizes its energy, the energy of the entire atom is a minimum as 
well. 

Review Questions 

1. While we have talked about emission spectra, another type of spectra is known as absorption spectra. 
In emission spectra, the atom emits lines of light like those you saw in the examples of atomic spectra. 
In absorption spectra, the atom absorbs lines of light, rather than emitting them. Can you explain 
this in terms of electrons and orbitals? What do you think the relationship between absorption 
spectra and emission spectra might be? 

2. If an electron has a "choice" between going into an orbital in the n = 1 energy level or an orbital in 
the n = 2 energy level, which do you think it chooses? 

3. If an electron in the n = 3 energy level has a "choice" between going into an orbital with t = or an 
orbital with t = 1, which do you think it chooses? 

4. Select the correct statement. According to the Aufbau Principle 

(a) orbitals with higher values of n fill up first 

(b) orbitals in the same energy level, but with higher values of t fill up first 

(c) orbitals with lower values of n fill up first 

(d) it is impossible to predict which orbitals will fill up first 

5. Decide whether each of the following statements is true or false. 

(a) Electrons in different orbitals have different energies. 

(b) An electron will enter an orbital of higher energy when a lower energy orbital is already filled. 

(c) For some atoms the first energy level can contain more than two electrons. 

6. Does the electron in the hydrogen atom absorb or emit energy when it makes a transition between 
the following energy levels: 

(a) n = 2 to n = 4 

(b) n = 6 to n = 5 

(c) n = 3 to n = 6 

7. Fill in the blanks. There is one s orbital, three p orbitals, and five d orbitals in the n = 3 energy level 
of an atom. If a particular atom has a total of 5 electrons in the n = 3 energy level, then there are... 

(a) electrons in the s orbital 

(b) electrons in p orbitals 

(c) electrons in d orbitals 

8. Fill in the blanks. There is one s orbital, three p orbitals, five d orbitals and 7 / orbitals in the n = 4 
energy level of an atom. If a particular atom has a total of 7 electrons in the n = 4 energy level, then 
there are... 

(a) electrons in the s orbital 

239 www.ckl2.org 



(b) electrons in p orbitals 

(c) electrons in d orbitals 

(d) electrons in / orbitals 

9. According to the Aufbau rule, which of the following atoms has a sub-shell that is exactly half-filled? 

(a) Ba 

(b) Al 

(c) C 

(d) As 

(e) O 

Further Reading / Supplemental Links 

• http://www. iun.edu/~cpanhd/C101webnotes/modern-atomic-theory/aufbau-principle.html 

• http : //www . avogadro . co . uk/light/auf bau/auf bau . htm 

• http://en.wikipedia.org/wiki 

Vocabulary 

Aufbau principle Electrons will fill available orbitals starting with those at the lowest energy before 
moving to those at higher energies. 

7.4 Writing Electron Configurations 

Lesson Objectives 

• Figure out how many electrons can exist at any given sublevel. 

• Figure out how many different sublevels can exist at any given energy level. 

• Be able to write electron configuration of any element given the total number of electrons in that 
element. 

• Be able to write either orbital representations or electron configuration codes. 

Introduction 

It's pretty easy to say 'electrons will fill the lowest available energy orbital'. Figuring out what that lowest 
available energy orbital is, though, can be quite a challenge. In this lesson, we're going to try to decide 
exactly which orbitals get filled, and when. Most people who are just beginning to learn about quantum 
chemistry tend to find orbital filling problems very confusing, so we'll take it slowly, and show you how 
diagrams and rules can help you to remember the ways in which different orbitals are organized and filled 

Electron Configurations 

How do electrons fill the different energy levels, energy sublevels, and orbitals? To understand that question, 
it helps a lot to look at a diagram. Figure 1 shows the first three energy levels (marked by the large 
differently colored blocks), and the sublevels (separated by dotted lines) that are present in any atom. In 

www.ckl2.org 240 



Figure 1, each of the circles represents an orbital and, of course, each can hold a total of two electrons. 
Notice the red n = 1 block contains only an s orbital. That orbital will hold the first two electrons in the 
atom. Once the red n = \ block is entirely filled, electrons will start filling the orange n = 2 block. In the 
orange n = 2 block, there are four different orbitals. The first orbital is an s orbital and the other three 
are p orbitals. Once all four orbitals in the orange n = 2 block have been filled, electrons will start filling 
the yellow n = 3 block. In the yellow n = 3 block, there are nine different orbitals. The first orbital is an 
s orbital, the next three are p orbitals, and the last five are d orbitals. When filling the n = 3 block, the s 
sublevel will always be filled first, since it is lowest in energy. Again, it can hold at most 2 electrons. After 
that, the next six will fill the three p orbitals. 




Q 



OOO 



OOOOO 



Figure 7.3: These are the first three energy levels with their sub-levels. 

So far we've talked about filling the orbitals in Figure 7.3 up to the 3p orbitals. That's a total of 18 
electrons. The first 18 electrons are 'nice and easy', because they fill the orbitals in order. First, all of the 
n = 1 orbitals get filled, then all of the n = 2 orbitals get filled, then the n = 3 orbitals get filled. From 
the 19 rfi electron on, though, things get a little crazy! Before we move on to the that region beyond 18 
electrons, let's take a brief look at a shorthand notation that scientists use to signify the electron orbital 
filling of a given atom. 

Since orbitals are filled in order of increasing n and, within each energy level, in order of increasing £, 
scientists can use a short hand, known as the electron configuration code, to represent filled orbitals. To 
write the electron configuration code for an atom, you write the symbol for the type of orbital present at a 
particular sublevel (Is, 2s, 2p, etc.) followed by a superscript to indicate how many electrons are actually 
in that sublevel in the atom you are describing. Let's take a look at a few examples. 

Example 1: 

Nitrogen has 7 electrons. Write the electron configuration for nitrogen. 

Take a close look at Figure 1, and use it to figure out how many electrons go into each sublevel, and also 
the order in which the different sublevels get filled. 

1. Begin by filling up the Is sublevel. This gives Is 2 . Now all of the orbitals in the red n = 1 block 

are filled. 
Since we used 2 electrons, there are 7-2 = 5 electrons left 

2. Next, fill the 2s sublevel. This gives ls 2 2s 2 . Now all of the orbitals in the s sublevel of the orange 

n = 2 block are filled. 
Since we used another 2 electrons, there are 5-2 = 3 electrons left 

3. Notice that we haven't filled the entire n = 2 block yet... there are still the p orbitals! The final 3 

electrons go into the 2p sublevel. This gives ls 2 2s 2 2p^ 

241 www.ckl2.org 



The overall electron configuration is: ls 2 2s 2 2p^ . 

Overlapping Energy Levels 

If you had to guess, which orbital do you think would be filled after the 3p orbitals? Most likely you'd 
guess that the 3d orbitals would come next - and that guess makes a lot of sense. Unfortunately, that 
guess is also wrong! It turns out that the 4s orbitals are filled before the 3d orbitals, even though the 
4s orbitals have n = 4 and the 3d orbitals only have n = 3. 

How is that possible? Does it mean that electrons go into higher energy orbitals before completely filling 
the lower energy orbitals? Is there something wrong with those electrons? Are certain electrons prevented 
from entering low energy orbitals? What's going on? 

It turns out that there's nothing wrong with those 4s electrons at all. They're still behaving like normal 
electrons and they're still going into the lowest energy orbital available. The only difference is that the 4s 
orbital is a lower in energy than the 3d orbitals. Sometimes, we get tricked into thinking that the 
principal quantum number determines which orbitals will get filled first. When it comes to the order in 
which orbitals are filled, though, the principal quantum number isn't the only factor - what also matters 
is the energy of the orbital. Usually orbitals with lower principal quantum numbers have lower energies, 
but that isn't the case when you compare the energies of the 3d orbitals and the energies of the 4s orbitals. 
In this case, the 4s orbitals are lower in energy than the 3d orbitals even though they have higher principal 
quantum numbers. Figure 7.4 shows a modified version of Figure 7.3, which shows the n = 4 orbital 
positions. Let's take a look at an example of an electron where some n = 4 orbitals are filled. 



» OOOOO- -* d 
O 5s 

OOO 4 p 

m OOOOO y 

(J 4s 

o ooo ; 

o QQQ 2 



A 



-Is 



Increasing 
Energy 



Figure 7.4: Notice that the orbital from the next higher energy level has slightly lower energy than the 
orbitals in the lower energy level. 

Example 2: 

Potassium has 19 electrons. Write the electron configuration code for potassium. 
This time, take a close look at Figure 7.4. 

1. Begin by filling up the Is sublevel. This gives Is 2 . Now the n = 1 level is filled. Since we used 2 

electrons, there are 19 - 2 = 17 electrons left 

2. Next, fill the 2s sublevel. This gives ls 2 2s 2 

Since we used another 2 electrons, there are 17 - 2 = 15 electrons left 

3. Next, fill the 2p sublevel. This gives ls 2 2s 2 2p e . Now the n = 2 level is filled. Since we used another 

6 electrons, there are 15-6 = 9 electrons left 

www.ckl2.org 242 



4. Next, fill the 3s sublevel. This gives ls 2 2s 2 2p 6 3s 2 

Since we used another 2 electrons, there are 9-2 = 7 electrons left 

5. Next, fill the 3p sublevel. This gives ls 2 2s 2 2p 6 3s 2 3p 6 

Since we used another 6 electrons, there are 7-6=1 electron left 

Here's where we have to be careful — right after 3/? 6 !! 
Remember, 4s comes before 3d\ 

6. The final electron goes into the As sublevel. This gives ls 2 2s 2 2p e 3s 2 3p e As 1 
The overall electron configuration code is: ls 2 2s 2 2p e 3s 2 3p e 4:S 1 . 

The Diagonal Rule 

Unfortunately, 4s orbitals aren't the only ones that get filled earlier than you'd expect based on their 
principal quantum number. The same turns out to be true of the 5s orbitals as well. Even though 5s 
orbitals have a higher principal quantum number than Ad orbitals, (n = 5 compared to n = 4), they're 
actually lower in energy. As a result, 5s orbitals are always filled before 4d orbitals. Similarly, 6s orbitals 
are lower in energy than 5d orbitals, so 6s orbitals are always filled first. The story gets even stranger 
when you consider / orbitals. 5s, 5p and 6s orbitals are all lower than 4/ orbitals. In other words, before 
you can get an electron into a 4/ orbital, you must first fill up the 5s orbitals, and the 5p orbitals and the 
6s orbitals! 

Filling up orbitals and writing electron configurations was so easy for atoms with less than 18 electrons! 
But for atoms with more than 18 electrons, it seems hopeless to memorize all of the different rules. How 
will you ever get straight whether the 5s orbital is higher or lower than the 4d orbital or the 4/ orbital? 
Thankfully, there's a simple rule known as the diagonal rule. The diagonal rule states that: Electrons fill 
orbitals in order of increasing 'quantum number sum' (n + €). When two orbitals share the same quantum 
number sum, they will be filled in order of increasing n. 

Luckily, the diagonal rule also has a diagram that's easy to remember, and that allows you to easily figure 
out the order in which electron orbitals are filled. Figure ?? shows the diagram. In order to use it, you 
must follow the arrows from tail-to-tip, starting with the first arrow in the upper left-hand corner, and 
working your way down through the arrows to the lower right-hand corner of the diagram. To see how this 
works, let's take a look at an example. 




Example 3: 



243 



www.ckl2.org 



Hafnium has 72 electrons. Write the electron configuration for hafnium. 

We can follow the arrows in the diagram (as shown below), until we have assigned all 72 electrons. Notice 
how you fill orbitals as you progress along the arrows, starting from the top arrow and moving down. 
Remember to stop once you hit 72. At that point, you have finished writing the electron configuration for 
Hafnium. 



(total: 2 electrons) 1s 
(total: 4 electrons) 1s 2s 2 



(total: 12 electrons) 1s 2 2s2p3s 2 




(total: 20 electrons) 1s 2 2s2p3s3p4s 

(total: 38 electrons) Is^p^s^pVSdVss 2 

(total: 56 electrons) 1s 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4cl5p 6 6s 2 

(total: 72 electrons) 1s 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d ,0 4p 6 5s4d5p6s4l ?4 5d 2 



Now that we know how electrons are assigned to orbitals, we can start to talk about how different orbitals, 
and different electron configurations actually affect the chemical properties of different atoms. In other 
words, we can actually start to talk about chemistry! Sometimes it seems that quantum physics and 
quantum chemistry are very far removed from the real world chemistry that we see and use in every 
day life. How are orbitals important when it comes to developing new drugs to treat cancer? How are 
energy levels important when it comes to inventing different kinds of superconductors or plastics? How are 
electron standing waves important when it comes to testing for toxins in your food? All of these processes 
are determined by chemical properties which are, themselves, a direct result of how electrons are arranged, 
and interact within different atoms and molecules. Sometimes it is easy to forget about the electrons, and 
some of the strange quantum properties of subatomic particles. Nevertheless, we will never be able to 
fully understand chemistry, if we don't understand its smallest components - components like electrons, 
protons, neutrons and, of course, the atom. 



Lesson Summary 

• For any atom with less than 18 electrons, orbitals are filled in order of increasing n and, for any given 
n, in order of increasing I. 

• Electron configurations are a shorthand notation for representing the filled orbitals in a given atom. 
They are written using the principal quantum number, n, for the energy level, the letter (s,p,d or /) 
for the sublevel, and a superscript for the number of electrons in that sublevel. 

• For atoms with more than 18 electrons, the orbitals are filled in order of increasing n (and in order 
of increasing I for a given n) up to the 18 electron; however after the 18 f ' ! electron the 4s orbital are 
filled before the 3d orbitals. This is because the 4s orbital has lower energy than the 3d orbital. 

• The diagonal rule states that electrons fill orbitals in order of increasing 'quantum number sum' 
(n + €). When two orbitals share the same 'quantum number sum,' they will be filled in order of 
increasing n. 



www.ckl2.org 



244 



Review Questions 

1. Write the electron configuration for beryllium. Beryllium has 4 electrons. 

2. Write the electron configuration for silicon. Silicon has 4 electrons. 

3. Write the electron configuration for nitrogen. Nitrogen has 7 electrons. 

4. Write the electron configuration for chromium. Chromium has 24 electrons. 

5. Write the electron configuration for silver. Silver has 47 electrons. 

Further Reading / Supplemental Links 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 3-1 is on the Development of the Atomic Theory. 

• Lesson 3-2 is on the Development of the Atomic Model. 

• Lesson 3-3 is on Atomic Structure. 

• Lesson 3-4 is on the Periodic Table. 

• Lesson 3.6 is on Electron Configuration, http : //www . f ordhamprep . org/gcurran/sho/sho/lessons/ 
lesson31 .htm 

Vocabulary 

electron configuration A short hand notation to indicate the electron orbitals which are filled in a 
particular atom. 

diagonal rule The electrons fill orbitals in order of increasing 'quantum number sum' {n + l). When two 
orbitals share the same 'quantum number sum', they will be filled in order of increasing n. 

quantum number sum The sum of the principal quantum number, n, and the azimuthal quantum 
number, t, for an electron. That is n + I. 



Image Sources 



(1) http: //en. wikipedia.org/wiki/File: Wolf gang_Pauli_young.jpg. Public Domain. 

(2) http://en.wikipedia.0rg/wiki/Image:UhlenbeckKramersG0udsmit.jpg. Public Domain. 

(3) Richard Parsons. . CC-BY-SA. 

(4) Richard Parsons. . CC-BY-SA. 



245 www.cki2.0rg 



Chapter 8 

Electron Configurations and 
the Periodic Table 



8.1 Electron Configurations of Main Group Ele- 
ments 

Lesson Objectives 

• Explain how the elements in the Periodic Table are organized into rows and columns. 

• Explain how the electron configurations within a column are similar to each other. 



Introduction 



It probably seems like all we've been spending a lot of time learning about protons... and neutrons... and 
electrons... and electrons... and more electrons... so you might be wondering - when do we actually get to 
study chemistry? When do we get to study reactions? When do we get to study explosions? When do 
we get to study plastics, and medicines that can be made by combining different kinds of chemicals? The 
answer is now. We're finally ready to discuss the chemical properties of the simplest chemicals out there 
- we're finally ready to discuss the elements. Remember, you have learned that there were 117 different 
kinds of atoms, and that each was known as an element. And you have learned that atoms of different 
elements have different numbers of protons. Hydrogen has 1 proton (and 1 electron if it's neutral), helium 
has 2 protons (and 2 electrons, if it's neutral), and lithium has 3 protons (and 3 electrons, if it's neutral). 
Finally, you have seen examples of the Periodic Table. Scientists use the Periodic Table to summarize 
information about all of the known elements that exist in our world. 

In this lesson, you will learn why the Periodic Table (as shown below) has such an unusual shape. 

(Source: CK-12 Foundation. CC-BY-SA) 



www.ckl2.org 246 



1 


































18 


1A 


2 


















13 


14 15 16 


17 


8A 


i 

H 


2 

He 


HYDROGEN 


2A 


















3A 


4A 5A 6A 


7A 


HELIUM 


3 

Li 


4 

Be 


5 

B 


6 

c 


7 

N 


8 




9 

F 


10 

Ne 


LITHIUM 


BERYLLIUM 


3 


4 


5 


6 


7 


8 9 10 


ii 


12 


BOHON 


CARBON 


NITROGEN 


OXYGEN 


FLUORINE 


NEON 


Na 


MAGNESIUM 


13 

Al 


14 

Si 


15 

P 


16 

s 


17 

CI 


18 

Ar 


SODIUM 


3B 


4B 


5B 


6B 


7B 


1 SB 1 


IB 


2B 


ALUMINUM 


SILICON 


PHOSPHORUS 


SULFUR 


CHLORINE 


ARGON 


19 

K 


20 

Ca 


21 

Sc 


22 

Ti 


23 

V 


24 

Cr 


25 

Mn 


26 

Fe 


27 

Co 


28 

Ni 


29 

Cu 


30 

Zn 


31 

Ga 


32 

Ge 


33 

As 


34 

Se 


35 

Br 


36 

Kr 


POTASSIUM 


CALCIUM 


SCANDIUM 


TITANIUM 


VANADIUM 


CHROMIUM 


MANGANESE 


IRON 


COBALT 


NICKEL 


COPPER 


ZINC 


GALLIUM 


GERMANIUM 


ARSENIC 


SELENIUM 


BROMIUM 


KRYPTON 


33 

Rb 


38 

Sr 


39 

Y 


40 

Zr 


41 

Nb 


42 

Mo 


43 

Tc 


44 

Ru 


45 

Rh 


46 

Pd 


47 


48 

Cd 


49 

In 


50 

Sn 


51 

Sb 


52 

Te 


53 

1 


54 

Xe 


RUBIDIUM 


Strontium 


YTTRIUM 


ZIRCONIUM 


NIORIUM 


MPLYRDENUM 


TECHNETIUM 


RUTHENIUM 


RHDDIUM 


PALLADIUM 


SILVER 


CADMIUM 


INDIUM 


TIN 


ANTIMONY 


TELLURIUM 


IODINE 


XENON 


55 

Cs 


56 

Ba 


57-71 

La-Lu 


72 

Hf 


73 

Ta 


74 

w 


75 

Re 


76 

Os 


77 

lr 


78 

Pt 


79 

Au 


80 


81 

TI 


82 

Pb 


83 

Bi 


84 

Po 


85 

At 


86 

Rn 


CESIUM 


BARIUM 


LANTHANIOES 


HAFNIUM 


TANTALUM 


TUNGSTEN 


RHENIUM 


OSMIUM 


IRIDIUM 


FLATIUM 


GOLD 


MERCURY 


THALUUM 


LEAD 


BISMUTH 


POLONIUM 


ASTATINE 


RADDN 


87 

Fr 


88 

Ra 


89-103 

Ac-Lr 


104 

Rf 


105 

Db 


106 

sg 


107 

Bh 


108 

Hs 


109 

Mt 


110 

Ds 


111 

Rg 


112 

Cn 


113 

Uut 


114 

Uuq 


115 

Uup 


116 

Uuh 


117 

Uus 


118 

Uuo 


FRANCIUM 


RADIUM 




RUTHERFORDIUM 


DUBNIUM 


SEABORGIUM 


BDHRIUM 


HASSIUM 


MEITNERIUM 


,_,«,.•„,. 




cor-ER-iciu... 


UNUNTHIUM 


UNUNQUADIUM 


UNUNPENTIUM 


_«!_ 


.»u„™. 


UHUHOCTIUM 



LANTHANIDES 


57 

La 


58 

Ce 


59 

Pr 


60 

Nd 


61 

Pm 


62 

Sm 


63 

Eu 


64 

Gd 


65 

Tb 


66 


67 

Ho 


68 

Er 


69 

Tm 


70 

Yb 


71 

Lu 




LANTHANUM 


CERIUM 


MOTI-M 


NEODYMIUM 


PROMETHIUM 


SAMARIUM 


EUROPIUM 


GADOLINIUM 


TERBIUM 


BVSPROSIUM 


HDLMIUM 


ERBIUM 


THULIUM 


YTTERBIUM 


LUTETIUM 


ACTINIDES 


89 

Ac 


90 

Th 


91 

Pa 


92 

u 


93 

Np 


94 

Pu 


95 

Am 


96 

Cm 


93 

Bk 


98 

Cf 


99 

Es 


100 

Fm 


101 

Md 


102 

No 


103 

Lr 




ACTINIUM 


THORIUM 


PROTACTINIUM 


URANIUM 


NEPTUNIUM 


PLUTONIUM 


AMERICIUM 


CURIUM 


BERKELIUM 


CALIFORNIUM 


EINSTEINIUM 


.EPMIUM 


MENOELEVIUM 


N OBELI UM 


•w 



Now, what's the first thing you thought when you saw the Periodic Table? If you're like most people, the 
first thing you thought was probably something like, 'Wow - that's a funny shape! Why is the Periodic 
Table shaped like that? Why is it lower in the middle? Why is it higher on either end? Why is there 
that odd- looking disconnected piece at the bottom? The Periodic Table doesn't look like a table at all!' 
In this chapter, you'll begin to see why the Periodic Table has such a funny shape. It turns out that the 
shape of the Periodic Table actually helps to tell us about the chemical properties of the different elements 
that exist in our world. In this section, for example, you'll learn that elements in the same column of the 
Periodic Table have similar chemical properties. Later we'll take a look at how elements in the same row 
are related. 

Group 1A Elements Have One s Electron 

Remember that according to the Aufbau principle electrons are added to low energy orbitals first and then, 
as the low energy orbitals are filled up, electrons go into higher and higher energy orbitals. When one 
atom reacts with another atom in a chemical reaction, it's the high-energy electrons that are involved. 

Since it's only the high-energy electrons that participate in a chemical reaction, it's only the high-energy 
electrons that we will concern us when we want to determine the chemical properties of a particular element. 
Just how 'high' in energy does an electron need to be to participate in a chemical reaction? Well, in most 
chemical reactions, the only electrons involved are the electrons in the highest energy level. In other words, 
the electrons with the highest value of n (the principal quantum number), participate in chemical reactions, 
while the electrons with lower values of n are called "core electrons," are closer to the nucleus and, as a 
result, don't get involved. The electrons with the highest value of n are known as valence electrons. Core 
electrons are also referred as non-valence electrons. Two different elements have similar chemical 
properties when they have the same number of valence electrons in their outermost energy level. 

Elements in the same column of the Periodic Table have similar chemical properties. So what does that 
mean about their valence electrons? You guessed it! Elements in the same column of the Periodic Table 
have the same number of valence electrons - that's why they have similar chemical properties. Let's see if 



247 



www.ckl2.org 



this is true for some of the elements in the first column of the Periodic Table. 

Example 1: 

Write the electron configuration for hydrogen (//). 

First, you need to find hydrogen on the Periodic Table. Take a look at the Periodic Table above. You 
know that hydrogen is in the first column, and if you look carefully, you'll see that hydrogen also happens 
to be at the top of the first column. The Periodic Table tells you that the atomic number for hydrogen 
is Z = 1, thus hydrogen has 1 proton. Neutral hydrogen will also have 1 electron. You need to write the 
electron configuration for an atom with 1 electron. 

As shown in the figure below, the diagonal rule applied to hydrogen (H). 



(total: 1 electrons 



principal quantum 
number: n = 1 



1 valence electron in an s orbital 




Therefore, we write the electron configuration for H : Is 1 . 

Remember, when you write electron configurations, the number out in front always indicates the principal 
quantum number, n, of a particular orbital, thus Is 2 has n = 1, while 3s 1 has n = 3. What is the highest 
principal quantum number that you see in hydrogen's electron configuration? It's n = 1, so all electrons 
with n = 1 are valence electrons. Hydrogen has 1 valence electron in an s orbital. 

Example 2: 

Write the electron configuration for lithium (Li). 

First, you find lithium on the Periodic Table. The Periodic Table tells you that the atomic number for 
lithium is Z = 3, thus lithium has 3 protons. Neutral lithium will also have 3 electrons. You need to write 
the electron configuration for an atom with 3 electrons. 

As illustrated in the figure below, the diagonal rule applied to lithium (Li), non-valence electrons: Is 2 . 



(total: 3 electrons 



(total: 2 electrons) ls 2 -A 

lis 2 2s 1 *2*^^*V^^' 

principal quantum 
number: n = 2 




1 valence electron in an s orbital 




Therefore, we write the electron configuration for Li : ls 2 2s 1 . 

What is the highest principal quantum number that you see in lithium's electron configuration? 



www.ckl2.org 



248 



It's n = 2, so all electrons with n = 2 are valence electrons, and all electrons with n < 2 are non-valence 
electrons. Lithium has 1 valence electron in an s orbital. 

Example 3: 

Write the electron configuration for sodium (Na). 

First, you find sodium on the Periodic Table. The Periodic Table tells you that the atomic number for 
sodium is Z = 11, thus sodium has 11 protons. Neutral sodium will also have 11 electrons. You need to 
write the electron configuration for an atom with 11 electrons. 

As shown below, the diagonal rule applied to sodium (Na). non-valence electrons: ls 2 2s 2 2p e . 



(total: 2 electrons) Is 2 

(total: 4 electrons) Is 2 2s 2 

(total: 11 electrons) Is 2 2s 2 2p 6 s 1 





principal quantum 
number: n = 3 

1 valence electron in an s orbital 



Therefore, we write the electron configuration for Na : ls 2 2s 2 2p e 3s 1 . 

What is the highest principal quantum number that you see in sodium's electron configuration? 

It's n = 3, so all electrons with n = 3 are valence electrons, and all electrons with n < 3 are non-valence 
electrons. (Don't be fooled by the 2p 6 orbitals. Even though they are p orbitals, not s orbitals, they have 
n = 2, so they are non-valence electrons!) Sodium has 1 valence electron in an s orbital. 

If you look at the last line in Example 1, Example 2, Example 3 you should notice a pattern. 

Hydrogen has 1 valence electron in an s orbital 

Lithium has 1 valence electron in an s orbital 

Sodium has 1 valence electron in an s orbital 

In fact, all elements in the first column of the Periodic Table have 1 valence electron in an s orbital. 

Therefore, we would expect all of these elements to have similar chemical properties - and they do. 
(Hydrogen is special because it is the first element in the Periodic Table. As a result, hydrogen has only 
one proton and one electron, which give it special chemical properties. Sometimes scientists don't include 
hydrogen in the first column of the Periodic Table, but instead give it its own 'special' column to reflect 
its special properties - we won't do that here, but you should realize that hydrogen does not have all the 
same chemical properties as the rest of the elements in its column.) 

The elements in the first column of the Periodic Table (other than hydrogen) are known as Group 1A 
metals, or Alkali Metals. When you compare the chemical properties of these elements (lithium, sodium, 
potassium, rubidium, cesium, and francium), what you'll notice is that they are all remarkably similar. 
Group 1A elements are metals, silver-colored, and soft. These elements are extremely reactive. Several of 
them explode if you put them in water. 

As pictured below, notice how the elements lithium (Li), sodium (Na), and potassium (K) all look alike. 
They are all soft, silver metals. Since Li,Na and K are all Group 1A metals, they all share similar chemical 
properties. (Source: http://en.wikipedia.Org/wiki/File:Limetal.JPG, http://en.wikipedia.org/ 
wiki/File : Nametal . JPG . jpg, http : //en . wikipedia . org/wiki /Image : Kmetal . jpg . CC-BY-SA) 



249 



www.cki2.0rg 




And finally, because they are so reactive, Group 1A elements are not found in their elemental form in 
nature - in other words, you don't find pure sodium or pure potassium in nature. 

As shown below, the diagonal rule applied to rubidium (Rb). non- valence electrons: ls 2 2s 2 2p^3s 2 3p^As 2 3d w Ap^. 



(total: 2 electrons) Is 2 — 

Is 2 2s 2 -** rs -^-^^ ^^' 



(total: 4 electrons) 
(total: 12electrons)ls 2 2s 2 2 
(total: 20electrons)ls 2 2s 2 2p 6 3s 2 3p 6 4s 
(total: 37electrons)ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 s 1 



principal quanturr 
number: n = 5 







1 valence electron in an s orbital 



Group 2A Elements Have Two s Electrons 



All of the elements in the first column of the periodic table have 1 valence electron in an s sublevel. How 
do you think the elements in the second column of the periodic table differ? Let's find out by taking a 
look at a few examples. 

Example 4: 

Write the electron configuration for beryllium (Be). 

First, you find beryllium on the Periodic Table. The Periodic Table tells you that the atomic number for 
beryllium is Z = 4, thus beryllium has 4 protons. Neutral beryllium will also have 4 electrons. You need 
to write the electron configuration for an atom with 4 electrons. 

Therefore, we write the electron configuration for Be : ls 2 2s 2 . 

What is the highest principal quantum number that you see in beryllium's electron configuration? 

It's n = 2, so all electrons with n = 2 are valence electrons, and all electrons with n < 2 are non-valence 
electrons. Beryllium has 2 valence electrons in an s orbital. 

As shown below, the diagonal rule applied to beryllium (Be), non- valence electrons: Is 2 . 



www.ckl2.org 



250 



(total: 2 electrons) is 2 
(total: 3 electrons) Is 2 2s 2 



principal quantum 
number: n = 2 



2 valence electron in an s orbital 





Example 5: 

Write the electron configuration for magnesium (Mg). 

First, you find magnesium on the Periodic Table. The Periodic Table tells you that the atomic number 
for magnesium is Z = 12, thus magnesium has 12 protons. Neutral magnesium will also have 12 electrons. 
You need to write the electron configuration for an atom with 12 electrons. 

Therefore, the electron configuration for Mg : ls 2 2s 2 2p 6 3s 2 . 

What is the highest principal quantum number that you see in strontium's electron configuration? 

It's n = 3, so all electrons with n = 3 are valence electrons, and all electrons with n < 3 are non-valence 
electrons. Magnesium has 2 valence electrons in an s orbital. 

Notice that: 

Beryllium has 2 valence electrons in an s orbital. 

Magnesium has 2 valence electrons in an s orbital. 

As pictured below, sports drinks frequently contain the electrolyte calcium chloride. (Source: CK-12 
Foundation. CC-BY-SA) 




You can probably guess the number and type of valence electrons in an atom of calcium (Co) , strontium 
(Sr), barium (Ba) or radium (Ra). If you guessed 2 electrons in an s orbital, then you guessed right! All 



251 



www.ckl2.org 



elements in the second column of the Periodic Table have 2 valence electrons in an s orbital. 

The elements in the second column of the Periodic Table are known as Group 2A metals, or Alkaline 
Earth Metals. As you might expect, because all Group 2A metals have 2 valence electrons in an s 
orbital, they all share similar chemical properties. Group 2A elements are metals, silver colored and are 
quite reactive though they are not nearly as reactive as the Group 1A elements. 

Group 3A Elements Have s and lp Electrons 

All of the elements in the first column of the Periodic Table have 1 valence electron in an s sublevel and all 
of the elements in the second column of the Periodic Table have 2 valence electrons in an s sublevel. Can 
you make any prediction about the valence electrons in the third column of the Periodic Table? Where 
is the third column of the Periodic Table? It turns out that there are really two different 'third columns' 
in the Periodic Table. Take a close look at the figure of the Period Table (the first figure of this lesson). 
Can you spot the column labeled '3A'? Can you spot the column labeled '3B'? Notice that the smallest 
atom in the '3B' column has Z = 21 (Scandium, Sc), while the smallest atom in the '3A' column has Z = 5 
(Boron, B). [You need to note that there is an alternate way to name 3A elements; the can also be referred 
to as group 13 since these elements are in the 13" 1 column of the Periodic Table.] Therefore, it obviously 
makes sense to discuss the 3A column first. Let's figure out how many valence electrons atoms in the 3A 
column have: 

Example 6: 

Write the electron configuration for boron (5). 

The Periodic Table tells you that the atomic number for boron is Z = 5, thus boron has 5 protons. Neutral 
boron will also have 5 electrons. You need to write the electron configuration for an atom with 5 electrons. 

Therefore, the electron configuration for B : ls 2 2s 2 2p 1 . 

What is the highest principal quantum number that you see in boron's electron configuration? 

It's n = 2, so all electrons with n = 2 are valence electrons, and all electrons with n < 2 are non-valence 
electrons. Both the electron in the 2p orbital and the electrons in the 2s orbital are valence electrons. 
Boron has 2 valence electrons in an s orbital, and 1 valence electron in a p orbital, for a 
total of 3 valence electrons. 

As pictured below, the diagonal rule applied to boron (B). 



(total: 2 electrons ) is 2 

(total: 3 electrons) Is 2 2s 2 

(total: 5 electrons) Is 2 2s 2 ^2p 1 ' 

principal quantum, 
number: n = 2 



3 valence electrons: 2 in an s orbital, 
1 in a p orbital 





Example 7: 

Write the electron configuration for aluminum (A/). 

The Periodic Table tells you that the atomic number for aluminum is Z = 13; thus neutral aluminum has 
13 protons and 13 electrons. You need to write the electron configuration for an atom with 13 electrons. 



www.ckl2.org 



252 



Therefore, the electron configuration for Al : ls 2 2s 2 2p (i 3s 2 3p 1 . 

What is the highest principal quantum number that you see in aluminum's electron configuration? 

It's n = 3, so all electrons with n = 3 are valence electrons, and all electrons with n < 3 are non-valence 
electrons. Both the electron in the 2>p orbital and the electrons in the 3s orbital are valence electrons. 
Aluminum has 2 valence electrons in an s orbital, and 1 valence electron in a p orbital, for 
a total of 3 valence electrons. 

From Example 7 and Example 8, we have: 

Boron has 2 valence electrons in an s orbital and 1 valence electron in a p orbital 

Aluminum has 2 valence electrons in an s orbital and 1 valence electron in a p orbital 

In fact, all elements in the 3A column of the Periodic Table have 2 valence electrons in an s orbital 
and 1 valence electron in a p orbital. That's a total of 3 valence electrons for atoms in the 3A column. 
Again, the chemical properties of 3A elements are similar, because they have the same number and type 
of valence electrons. 



Group 4A-8A Continue to Add p Electrons to the Outermost 
Energy Level 



By now, you may have noticed a pattern relating the number of valence electrons to the column number. 
Group 1A elements have 1 valence electron. Group 2A elements have 2 valence electrons. Group 3A 
elements have 3 valence electrons. Group 4A elements have... well, we haven't looked at them yet, but 
what would you guess? It's pretty obvious. Group 4A elements have 4 valence electrons. Similarly, Group 
5 A elements have 5 valence electrons. In fact, the pattern continues all the way up to Group 8 A elements, 
which have 8 valence electrons. Let's take a look at a few examples in order to figure out exactly what 
types of valence electrons are involved. First, we'll consider a Group 4A element. 

Example 8: 

Write the electron configuration for carbon (C). 

The Periodic Table tells you that the atomic number for carbon is Z = 6, thus neutral carbon has 6 protons 
and 6 electrons. You need to write the electron configuration for an atom with 6 electrons. 

Therefore, the electron configuration for C : ls 2 2s 2 2p 2 . 

What is the highest principal quantum number that you see in carbon's electron configuration? 

It's n = 2, so all electrons with n = 2 are valence electrons, and all electrons with n < 2 are non-valence 
electrons. Both the electrons in the 2p orbitals and the electrons in the 2s orbital are valence electrons. 
Carbon has 2 valence electrons in an s orbital, and 2 valences electron in p orbitals, for a 
total of 4 valence electrons. 

Illustrated below, the diagonal rule applied to carbon (C). non-valence electrons: Is 2 . 

253 www.ckl2.org 



(total: 2 electrons) Is 

(total: 4 electrons) Is 2 2s 2 

(total: 12 electrons) Is 2 s 2 p 

principal quantum 
number: n = 2 



4 valence electrons: 2 in an s orbital, 
and 2 in p orbitals 





Now let's consider a Group 5 A element. 

Example 9: 

Write the electron configuration for nitrogen (N). 

The Periodic Table tells you that the atomic number for nitrogen is Z = 7, neutral nitrogen has 7 protons 
and 7 electrons. You need to write the electron configuration for an atom with 7 electrons. 

Therefore, the electron configuration for N : ls 2 2s 2 2p 3 . 

What is the highest principal quantum number that you see in phosphorus's electron configuration? It's 
n = 2, so all electrons with n = 2 are valence electrons, and all electrons with n < 2 are non- valence 
electrons. Both the electrons in the 2p orbitals and the electrons in the 2s orbital are valence electrons. 
Nitrogen has 2 valence electrons in an s orbital, and 3 valence electrons in p orbitals for a 
total of 5 valence electrons. 

As a final example, let's take a look at a Group 6A element (or Group 16). 

Example 10: 

Write the electron configuration for oxygen (0). 

The Periodic Table tells you that the atomic number for oxygen is Z = 8; neutral oxygen has 8 protons 
and 8 electrons. You need to write the electron configuration for an atom with 8 electrons. 

Therefore, the electron configuration for O : ls 2 2s 2 2/A 

What is the highest principal quantum number that you see in oxygen 's electron configuration? 

It's n = 2, so all electrons with n = 2 are valence electrons, and all electrons with n < 2 are non-valence 
electrons. Both the electrons in the 2p orbitals and the electrons in the 2s orbital are valence electrons. 
Oxygen has 2 valence electrons in an s orbital, and 4 valence electrons in p orbitals, for a 
total of 6 valence electrons. 

So let's summarize what we know so far: 

Group 1A elements have 1 valence electron in an s orbital 

Group 2A elements have 2 valence electrons in an s orbital 

Group 3A elements have 2 valence electrons in an s orbital and 1 valence electron in a p orbital 

Group 4A elements have 2 valence electrons in an s orbital and 2 valence electrons in p orbitals 

Group 5A elements have 2 valence electrons in an s orbital and 3 valence electrons in p orbitals 

Group 6A elements have 2 valence electrons in an s orbital and 4 valence electrons in p orbitals 

Can you guess how the valence electrons in columns 7A (Group 17) and 8A (Group 18) are arranged? 



www.ckl2.org 



254 



Group 7A elements have 2 valence electrons in an s orbital and 5 valence electrons in p orbitals 

Group 8A elements have 2 valence electrons in an s orbital and 6 valence electrons in p orbitals 

Notice that, after column 3A, each column one step further to the right has one additional valence p 
electron. Group 4A elements have one more valence p electron than Group 3A elements. Similarly, Group 
5 A elements have one more valence p electron than Group 4A elements. But what happens when you 
reach Group 8A elements? Why does the Periodic Table end at column 8A? Let's think about that 
carefully. Group 8A elements have 6 valence electrons in p orbitals. In the last chapter, you learned that 
the maximum number of p electrons at any energy level is 6. Therefore, there couldn't be a '9A' column, 
because a '9A' column would have Ip electrons in the valence energy level, which is impossible. 

The fact that Group 8A elements have completely filled valence s sublevel and p sublevel is important in 
terms of their chemical properties. Group 8A elements are called Noble Gases. 

They are all gases, and they are not very reactive at all. 

Lesson Summary 

• All known elements are organized into the Periodic Table in such a way that elements in the same 
column have similar chemical properties. 

• Only the highest energy electrons (valence electrons) are involved in chemical reactions. Therefore, it 
is only these high-energy electrons that are important in determining an elements chemical properties. 

• Two different elements are likely to have similar chemical properties when they have the same number 
of valence electrons. 

• Elements with the same number of valence electrons are found in the same column of the Periodic 
Table. 

• All elements in the first column of the Periodic Table have 1 valence electron in an s orbital. These 
elements are known as Group 1A metals or alkali metals. 

• All elements in the second column of the Periodic Table have 2 valence electrons in an s orbital. 
These elements are known as Group 2A metals or alkaline earth metals. 

• All elements in column 3A of the Periodic Table have 2 valence electrons in an s orbital and 1 valence 
electron in a p orbital. 

• All elements in column 4A of the Periodic Table have 2 valence electrons in an s orbital and 2 valence 
electrons in p orbitals. ..etc. 

• Column 8A has 2 valence electrons in an s orbital and 6 valence in p orbitals. Since any given energy 
level can have at most dp electrons, column 8A elements have a filled p sublevel. Therefore, they are 
inert (non-reactive), because they are unlikely to either gain or lose electrons. Group 8A elements 
are called Noble Gases. 



Review Questions 

1. Take a look at the Periodic Table. How would you describe it? Why do you think it has such a funny 
shape? 

2. Can you suggest how elements in the same column of the Periodic Table might be similar? 

3. Choose the correct statement. 

(a) Mg has only 1 valence electron in an s orbital 

(b) F has only 1 valence electron in an s orbital 

(c) O has only 1 valence electron in an s orbital 

(d) Kr has only 1 valence electron in an s orbital 

(e) Fr has only 1 valence electron in an s orbital 

255 www.ckl2.org 



4. Circle the appropriate element for each blank. 

(a) (Mg/N) has 2 valence electrons in an s orbital, and 3 valence electrons in 

p orbitals. 
(b) (As/B) has 2 valence electrons in an s orbital, and 3 valence electrons in 

p orbitals 
(c) (Cl/P/Li) has 2 valence electrons in an s orbital, and 5 valence electrons 

in p orbitals 
(d) {Al/Li/Na) has 1 valence electron in a p orbital 

5. Choose the correct statement. 

(a) Group 1A elements have a total of 3 valence electrons 

(b) Group 5A elements have a total of 2 valence electrons 

(c) Group 7A elements have a total of 4 valence electrons 

(d) Group 8A elements have a total of 8 valence electrons 

(e) Group 2A elements have a total of 5 valence electrons 

(f) Group 1A elements have a total of 3 valence electrons 

6. Fill in the blanks. 

(a) N has valence electrons in an s orbital 

(b) ,/V has valence electrons in p orbitals 

(c) N has a total of valence electrons 

(d) Ca has valence electrons in s orbitals 

(e) Ca has valence electrons in p orbitals 

(f) Ca has a total of valence electrons 

7. Decide whether each of the following statements is true or false. 

(a) K has 1 valence electron in an s orbital 

(b) Ge has 2 valence electrons in an s orbital 

(c) Se has 4 valence electrons in p orbitals 

(d) B has 3 valence electrons in p orbitals 

(e) F has 2 valence electrons in an s orbital, and 7 valence electrons in p orbitals 

(f) Ca has a total of 4 valence electrons 

8. Match the element to its valence electrons. 

(a) Sr - i. a total of 8 valence electrons 

(b) / - ii. a total of 2 valence electrons 

(c) Ne - iii. a total of 5 valence electrons 

(d) N - iv. a total of 7 valence electrons 

9. Fill in the blanks. 

(a) Ba has valence electron(s) in an s orbital, and valence electron(s) in p orbitals 

(b) Sn has valence electron(s) in an s orbital, and valence electron(s) in p orbitals 

(c) S has valence electron(s) in an s orbital, and valence electron(s) in p orbitals 

(d) Po has valence electron(s) in an s orbital, and valence electron(s) in p orbitals 

(e) Na has valence electron(s) in an s orbital, and valence electron(s) in p orbitals 

10. List all of the elements with exactly 2 valence electrons in p orbitals. 

11. An element has 2 valence electrons in an s orbital and 4 valence electrons in p orbitals. If the element 
is in the second row of the Periodic Table, which element is it? 

12. An element has 2 valence electrons in an s orbital and 6 valence electrons in p orbitals. If the element 
is in the same row as In, which element is it? 

www.ckl2.org 256 



Further Reading / Supplemental Links 

• http : //www . wou . edu/las/physci/ch412/perhist .htm 

• http : //www . aip . org/history/curie/periodic . htm 

• http: //web. buddyproject . org/web017/web017/history .html 

• http://www.dayah.com/periodic 

• http : //www . chemtutor . com/perich . htm 

Vocabulary 

Periodic Table Scientists use the Periodic Table to summarize what they know about the existing 
elements. Elements of similar size are found in the same row, while elements with similar chemical 
properties are found in the same column. 

chemical properties The ways in which an element reacts with another element or compound. 

valence electrons The electrons in an atom with the highest value of n (the electrons in the highest 
energy level). 

no n- valence electrons All electrons in atom which are not valence electrons. Non- valence electrons are 
not important in determining an element's chemical properties because they rarely get involved in 
chemical reactions. 

alkali metals Group 1A metals. These are elements found in the first column of the Periodic Table, 
excluding hydrogen. 

alkaline earth metals Group 2A metals. These are elements found in the second column of the Periodic 
Table. 

noble gases Group 8A elements. These are elements found in the eight column of the Periodic Table. 
They are inert, which means that they are very non-reactive. 

8.2 Orbital Configurations 

Lesson Objectives 

• Draw orbital diagrams. 

• Define Hund's Rule. 

• Use Hund's Rule to decide how electrons fill sublevels with more than one orbital. 

257 www.ckl2.org 



Introduction 

In the Electron Configurations of Main Group Elements lesson, you learned a little bit about valence 
electrons. You saw how the number and type of valence electrons are important in determining the 
chemical properties of a particular element. Group 1A metals were highly reactive, because they have a 
strong tendency to lose their single valence s electrons. Group 2A metals are reactive as well, but less 
so, because they had 2 valence s electrons. Finally, Group 8A elements were inert (not reactive at all), 
because they had completely filled valences and p sub levels, meaning they could neither lose nor gain 
electrons very easily. Now you might be wondering why we didn't talk much about the chemical properties 
of the elements in columns 4A-7A. It turns out that understanding the behavior of these elements requires 
a bit more information. Specifically, we need to know how the electrons fill up the p orbitals. Carbon, 
for instance, is a Group 4A element, so it has 2 valence s electrons, and 2 valence p electrons. Obviously, 
the 2 valence s electrons are paired together in the s orbital, but what about the 2 valence p electrons? 
Are the valence p electrons paired in a single p orbital, or are they each in their own p orbital (remember, 
there are a total of three p orbitals that the valence p electrons could be found in). What about nitrogen? 
Nitrogen is a Group 5A element, so it has 2 valence s electrons, and 3 valence p electrons. Again, the 2 
valence s electrons must be paired in the s orbital, but what about the 3 valence p electrons? Are two of 
them paired in a single p orbital, or do all three have their own p orbitals? 



Orbital Representation 

Before we discuss the order and manner in which the orbitals in a p sublevel are filled, we have to introduce a 
symbolic notation that scientists use to show orbital filling. Frequently, scientists will use boxes to represent 
orbitals. 

The figure below shows a set of boxes for orbitals up to the n = 2 energy level. The boxes are depicting 
electron orbitals for the n = 1 and n = 2 energy levels. Notice that for n = 1, there is a single s orbital, 
while for n = 2 there is one s orbital and three p orbitals. 



Is 2s 2p 

When a 'spin-up' electron is present in an orbital, scientists draw an upward pointing arrow in that orbital's 
box. This leads to what is known as an orbital diagram. For example, hydrogen has a single electron in 
the Is orbital. If that single electron were a spin-up, the orbital diagram for hydrogen would be: 

Illustrated below, the orbital diagram for hydrogen, assuming that hydrogen's single electron is 'spin-up' 
(m, = +1/2). 



t 



Is 2s 2p 

When a 'spin-down' electron is present in an orbital, scientists draw a downward pointing arrow in that 
orbital's box. Helium, for instance, has two electrons in the Is orbital. We know that one of these electrons 
must be 'spin-up' and the other must be 'spin-down,' so the orbital diagram for helium would be: 

As shown below, the orbital diagram for helium. Notice that when electrons are paired in an orbital, one 
of the electrons is 'spin- up,' while the other is 'spin-down.' 

www.ckl2.org 258 



1 

Is 



2s 



2p 



By comparing the first two orbital representation figures, it should be fairly obvious to you, which electrons 
are paired and which are unpaired. That is one of the advantages of an orbital diagram. An orbital diagram 
is a clear way of showing exactly how many paired and unpaired electrons there are in a particular atom's 
electronic configuration. 

Now you might be wondering about the unpaired electrons - how do you know whether to draw them 
as 'spin-up,' or 'spin-down.' Technically speaking, the first electron in a sublevel (Is, 2s, 2p, etc) could be 
either 'spin-up' or 'spin-down.' In other words, for hydrogen (Is 1 ), you could draw the arrow in the Is 
orbital box pointing either up or down. Similarly, for boron (ls 2 2s 2 2p 1 ), you could draw the arrow in the 
2p orbital box pointing either up or down. Once you've chosen the spin of the first electron in a sublevel, 
though, the spins of all of the other electrons in that sublevel depend on the spin you chose for the first. 
To avoid confusion, scientists 'always draw the first electron in an orbital as 'spin-up." If you stick with 
this rule, you'll never get into trouble. 

If you insist on breaking this rule, you might draw an orbital diagram that is incorrect, as shown in the 
figure below. By convention, scientists usually draw all unpaired electrons as 'spin-up.' This prevents them 
from drawing incorrect orbital diagrams like the one shown in (c). In the next section, you'll learn that 
the orbital diagram in (c) is incorrect because it does not obey Hund's Rule. 



I E 

Is 2s 



correct 



2p 



"■IS 

Is 



m 

2s 



2p 



correct 

(since this is the 1 electron in the p sublevel, 
it can be either 'spin up' or 'spin down') 



ffl 

Is 



BE 

2s 



2p 



incorrect 

nd 

(the spin of the 2 electron depends on the spin of the 
f'electron, even though they aren't in the same orbital! 



ffl 

Is 



BE 

2s 



2p 



correct 

(as long as you draw the//rst electron in each orbital as 
'spin-up', you will draw a correct orbital diagram ) 



Notice that by drawing the first electron in each orbital as 'spin-up,' all of the unpaired electrons in the 
sublevel have the same spin, even if they are in different orbitals! This is due to a principle known as 
Hund's Rule. We'll discuss Hund's Rule in more detail in the next section. 



259 



www.ckl2.org 



Orbitals in p Sublevel Fill Individually Before Pairing 



Previously we learned that the energy of an electron in any given orbital depends on the energy level of the 
orbital (which is determined by the principal quantum number, n) and the sublevel of the orbital (s,p,d, 
etc. which is determined by the azimuthal quantum number: €). We also learned that according to the 
Aufbau principle, electrons will fill the lowest energy orbitals first, and then move up to higher energy 
orbitals only after the lower energy orbitals are full. If you think carefully, though, you'll realize that 
there's still a problem. Certainly, Is orbitals should be filled before 2s orbitals, because the Is orbitals 
have a lower value of n, and thus a lower energy. Similarly, 2s orbitals should be filled before 2p orbitals, 
because 2s orbitals have a lower value of l{t = 0), and thus a lower energy. What about the three different 
2p orbitals? In what order do electrons fill the 2p orbitals? To answer this question, we need to turn to a 
principle known as Hund's Rule. Hund's Rule states that: (1) Every orbital in a sublevel is singly occupied 
before any orbital is doubly occupied. (2) All of the electrons in singly occupied orbitals have the same 
spin. 

According to the first rule, electrons will always occupy an empty orbital before they pair up. This should 
make sense given what you know about electrons. Electrons are negatively charged and, as a result, they 
repel each other. Since electron-electron repulsion raises the energy of the electrons involved, electrons 
tend to minimize repulsion (and thus minimize their energies) by occupying their own orbital, rather than 
sharing an orbital with another electron. Take a look at the figure below. 

no clcctron-clcetrcn repulsion 
equals lower energy 



11 si eh 





electron- electron repulsion 
equals higher energy 


1*1 


e a 








Is 


2s 
incorret 


2p 
t 







Is 2s 2p 

correct 

Notice how the two 2p electrons in the orbital diagram on the left are in separate orbitals, while the two 2p 
electrons in the orbital diagram on the right are sharing a single orbital. The orbital diagram on the left 
is the correct orbital diagram, because it obeys Hund's Rule, meaning that there is less electron-electron 
repulsion and, as a result, the electrons have lower energies (remember, electrons always minimize their 
energies). 

no elcetron-elcetroi repulsion 
equals lower energy 



m m Ltun 





electron- electron repulsion 
equals higher energy 


1*1 


E St 


Is 


2s 2p 
incorrect 



Is 2s 2p 

correct 

The figure above illustrating orbital diagrams for nitrogen is similar to the orbital diagram for carbon in 
the first figure. Notice how all three 2p electrons in the orbital diagram on the left are in separate orbitals, 
while two of the three 2p electrons in the diagram on the right are sharing a single orbital. The orbital 
diagram on the left is the correct orbital diagram, because it obeys Hund's Rule. Again, this means that 
there is less electron-electron repulsion and, as a result, the electrons have lower energies. 



no electron-electron repulsion 
equals lower energy 


ffl 

Is 


electron- electron repulsion 
equals higher energy 


' 




* BE i»i_t 


E 5> 


Is 2s 2p 


2s 2p 


correct 




incorrect 


www.ckl2.org 




260 



The figure above shows possible electron configurations for an atom with four 2p electrons. This time, two 
of the electrons have no choice - they must pair up. The other two electrons, however, could either pair 
up, as shown in the orbital diagram on the right, or occupy their own orbitals, as shown in the orbital 
diagram on the left. Which do you think is correct? Obviously, the orbital diagram on the left, because it 
minimizes electron-electron repulsion. The orbital diagram on the left is also the orbital diagram, which 
follows Hund's Rule, since all orbitals are singly occupied before any are doubly occupied. The orbital 
diagram on the right does not follow Hund's Rule, since the first two orbitals are doubly occupied before 
the third is singly occupied. 

While it's easy to understand why electrons would occupy empty orbitals before pairing up, it's a lot harder 
to understand why unpaired electrons in different orbitals must all have the same spin. Electron spins in 
different orbitals align (all point in the same direction), because spins, which are aligned have lower energy 
than spins which are not aligned. Notice that as long as you always draw the first electron in an 
orbital as 'spin-up' you will always draw spins which are aligned. Refer back to the figure illustrating 
unpaired electrons as 'spin-up.' Now that you know Hund's Rule, it should be obvious why the orbital 
diagram in 4.c is incorrect - the two electrons in the singly occupied 2p orbitals have different spins, and 
thus this orbital diagram does not obey Hund's Rule. Compare the orbital diagram in 4.c to the orbital 
diagram in 4. d. The orbital diagram in 4. d does obey Hund's Rule, because the two electrons in the singly 
occupied 2p orbitals have the same spin. 

Nitrogen Has Three Unpaired Electrons 

Now that we know about orbital diagrams and Hund's rule, we can begin to explain the chemistry of the 
elements in Groups 4A through 7A. Let's take a look at nitrogen as an example. 

Example 1: 

Draw the orbital diagram for nitrogen. 

First, we need to write the electron configuration for nitrogen just as we did previously: which gives 
ls 2 2s 2 2p' i . To draw the orbital diagram we will write the following: the first two electrons will pair up in 
the Is orbital; the next two electrons will pair up in the 2s orbital. That leaves 3 electrons, which must 
be placed in the 2p orbitals. According to Hund's Rule, all orbitals will be singly occupied before any 
is doubly occupied. Therefore, we know that each p orbital gets one electron. Hund's Rule also tells us 
that all of the unpaired electrons must have the same spin. Keeping with convention, we draw all of these 
electrons as 'spin-up,' which gives. 

The first two electrons pair The last three electrons singly occupy the three 

up in the Is orbital 2porbitals. They all have the same spin! 

\ J 

1 s] rrrrrr 



Is 2s 2p 

t 

The second two electrons 
pair up in the 2s orbital 

Orbital diagrams can help you to make predictions about the ways in which certain elements will react, 
and the chemical 'compounds' or 'molecules' that different elements will form. We aren't ready to discuss 
compounds, or molecules yet. The principles that you're learning now will help you to understand the 
behavior of all chemicals, from the most basic elements like hydrogen and helium, to the most complex 
proteins (huge biological chemicals made of thousands of different atoms bound together) found in your 
body. 

261 www.ckl2.org 



Lesson Summary 

• Orbital diagrams are drawn by representing each orbital as a box, each 'spin-up' electron in an orbital 
as an upward pointing arrow in the box, and each 'spin-down' electron in an orbital as a downward 
pointing arrow in the box. You can only have two arrows in each box, and they must be pointing in 
opposite directions. 

• Scientists use the convention that the first electron in any orbital is 'spin-up,' therefore, the first 
arrow in an orbital 'box' should point up. Hund's Rule states: 

(1) Every orbital in a sublevel is singly occupied before any orbital is doubly occupied. 

(2) All of the electrons in singly occupied orbitals have the same spin. 

• Electrons will occupy separate orbitals rather than pairing up, since this minimizes electron-electron 
repulsions, thereby minimizing energy. Electron spins in different orbitals within the same sublevel 
align because aligned spins have lower energy. 



Review Questions 

1. Which of the following is a valid orbital diagram? 



b. 



Tl 


Tl 

2s 

Tl 

2s 

Tl 

2s 

Tl 

2s 

u 


T 1 T 


Is 


2p 


Tl 


Tl 


Is 


2p 


U 


1 T T 


Is 


2p 


Tl 


Tl TT T 


Is 


2p 


Tl 


tl T T 



Is 



2s 



2p 



2. Draw the orbital diagram for lithium (Li). 

3. Draw the orbital diagram for carbon (C). 

4. Draw the orbital diagram for fluorine (F). 

5. Draw the orbital diagram for oxygen, O. Use it to answer the following questions: 



(a) an oxygen atom has 

(b) an oxygen atom has 

(c) an oxygen atom has 

(d) an oxygen atom has 



_ unpaired valence electrons 
_ paired valence electrons 
_ paired non-valence electrons 
_ unpaired non-valence electrons 

6. Draw the orbital diagram for neon, Ne. Use it to answer the following questions: 

(a) a neon atom has unpaired valence electrons 

(b) a neon atom has paired valence electrons 

(c) a neon atom has paired non-valence electrons 

(d) a neon atom has unpaired non- valence electrons 



www.ckl2.org 



262 



7. Decide whether each of the following statements is true or false. 

(a) Every orbital in a sublevel is doubly occupied before any orbital is singly occupied. 

(b) Every orbital in a sublevel is singly occupied before any orbital is doubly occupied. 

(c) All electrons in singly occupied orbitals have the same spin. 

(d) The two electrons in a single orbital have the same spin. 

(e) All electrons in singly occupied orbitals have different spins. 

(f) The two electrons in a single orbital have different spins. 

8. Draw the orbital diagram for phosphorus P. 

9. Draw an orbital diagram for silicon, Si. Use it to answer the following questions: 

(a) a silicon atom has unpaired valence electrons 

(b) a silicon atom has paired valence electrons 

(c) a silicon atom has paired non-valence electrons 

(d) a silicon atom has unpaired non-valence electrons 

10. Draw an orbital diagram for Mn. Use it to determine the total number of unpaired electrons in an 
Mn atom. 

Further Reading / Supplemental Links 

• http://www. chempractice. com/drills/ java_AO.php 

• http : //library . thinkquest . org/3659/structures/electronconf ig . html 

• http : //www . chem.latech.edu/~deddy/Lectnote/Chap7B . html 

• http://www.mi .mun.ca/users/edurnf or/1100/atomic°/ 20structure/sld036.htm 

Vocabulary 

orbital diagram Orbital diagrams are drawn by representing each orbital as a box, each 'spin-up' elec- 
tron in an orbital as an upward pointing arrow in the box, and each 'spin-down' electron in an orbital 
as a downward pointing arrow in the box. 

Hund's rule Every orbital in a sublevel is singly occupied before any orbital is doubly occupied. All of 
the electrons in singly occupied orbitals have the same spin. 

8.3 The Periodic Table and Electron Configura- 
tions 

Lesson Objectives 

• Relate an element's position in the PT to the energy level of its valence electrons (excluding transition 
metals, lanthanides, and actinides). 

• Relate an element's position in the PT to the sublevel of its highest energy valence electrons. 

• Explain why there are only two elements in the first row of the PT. 

263 www.ckl2.org 



Introduction 



With what we have already discussed, you might realize that just as electron configurations can be used 
to explain the shape and organization of the Periodic Table, the shape and organization of the Periodic 
Table can, in turn, be used to predict electron configurations. In fact, if you can locate an element on the 
Periodic Table, you can use the element's position to figure out the energy level of the element's valence 
electrons. Furthermore, an element's position on the Periodic Table tells you the sublevel of the element's 
highest energy valence electrons. In this lesson, we'll take a close look at how the Periodic Table relates to 
the electron configurations. 



Rows Across on the PT are Consistent With the Energy Level in 
An Atom 



First, let's try to figure out what we can learn from an element's row or period in the Periodic Table. 

The figure below shows how the different rows in the Periodic Table are numbered. The transition metals 
and the lanthanides and actinides have been omitted. (Source: CK-12 Foundation. CC-BY-SA) 



1 

1A 
















18 
8A 


l 

H 

[1.00784; 1.00811] 

HYDROGEN 


2 13 14 15 16 17 
2A 3A 4A 5A 6A 7A 


2 

He 

4.0026 

HEUUM 


3 

Li 

; [6.938; 6.997] 

LITHIUM 


4 

Be 

9.0122 

BERYLLIUM 




5 

B 

[10.S06; 10.821] 

BORON 


6 

c 

[12.0096; 12.0116] 

CARBON 


7 

JL 

NITROGEN 


8 



[15.99903:15.99977] 

OXYGEN 


9 

F 

FLUORINE 


10 

Ne 

20.180 

NEON 


Na 

22.990 

SODIUM 


12 

Mg 

MAGNESIUM 




13 

Al 

25.9B2 
ALUMINUM 


14 

Si 

[28.084; 28.086] 

SILICON 


15 

P 

30.974 

PHOSPHORUS 


16 

s 

[32.059; 32.075] 
SULFUR 


17 

CI 

[35.446; 35.457] 

CHLORINE 


18 

Ar 

39.94B 

ARGON 


19 

K 

39.09B 

POTASSIUM 


20 

Ca 

40.078 

CALCIUM 




31 

Ga 

69.723 

GALLIUM 


32 

Ge 

69.723 
GERMANIUM 


33 

As 

74.922 
ARSENIC 


34 

Se 

7B.963 

SELENIUM 


35 

Br 

79.904 
BR0MIUM 


36 

Kr 

83,801 
KRYPTON 


37 

Rb 

85.46S 

RUBIDIUM 


38 

Sr 

87.62 

STRONTIUM 




49 

In 

INDIUM 


50 

Sn 

TIN 


51 

Sb 

121.760 

ANTIMONY 


52 

Te 

127.603 

TELLURIUM 


53 

1 

126.904 

IODINE 


54 

Xe 

131.292 

XENON 




















55 

Cs 

132.905 

CESIUM 


56 

Ba 

137.327 

BARIUM 




81 

Tl 

[204.3B2; 204.385] 

THALLIUM 


82 

Pb 

204.383 

LEAD 


83 

Bi 

208.980 

BISMUTH 


84 

Po 

20B.98? 

POLONIUM 


85 

At 

209.9B7 

ASTATINE 


86 

Rn 

222.018 

RADON 


87 

Fr 

223.020 

FRANCIUM 


88 

Ra 

226.0254 

RADIUM 




113 

Uut 

284 

UNUNTRIUM 


114 

Uuq 

284 

UNUNQUAI1IUM 


115 

Uup 

288 
UNUNPENTIUM 


116 

Uuh 

292 

UNUNHFXIUM 


117 

Uus 

294 
UNUNSKPilUM 


118 

Uuo 

294 
UNUN0CTIUM 



ROW1 



ROW 2 



ROW 3 



ROW 4 



ROWS 



ROW 6 



ROW 7 



To understand what this means in terms of an element's electron configuration, let's consider the Group 
1A metals. If we write the electron configuration for the Group 1A metal from each row of the Periodic 
Table, we have: 



www.ckl2.org 



264 



row 
row 
row 
row 
row 
row 



2 
3 

4 

5 
6 

7 



Li: 

Na 

K: 

Rb 

Cs 

Fr 



l5 2 2/ 

ls 2 2s 2 2p 6 3s l 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 1 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 1 

ls 2 2s 2 2p 6 3s 2 3 P Hs 2 3d l0 4p 6 5s 2 4d l0 5p%s l 

s 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d 10 5p%s 2 4f u 5d w 6p 6 7s 1 



Do you see any pattern? For Group 1A metals, it seems that element's row corresponds to the energy level 
of that element's valence electron. Lithium {Li), for instance, is found in row 2 of the Periodic Table, and 
its valence electron is found in the n = 2 energy level. Cesium {Cs) is found in row 6 of the Periodic Table, 
and its valence electron is found in the n = 6 energy level. Let's see if this same pattern holds for Group 
2 A metals: 



row 
row 
row 
row 
row 
row 



2 
3 

4 

5 
6 

7 



Li: 

Na 
K: 
Rb 
Cs 
Fr 



ls 2 2s 2 

ls 2 2s 2 2p 6 3s 2 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 



ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 5p 6 6s 2 
ls 2 2s 2 2p 6 3s 2 3p (i 4s 2 3d 10 4p 6 5s 2 4d l0 5p (i 6s 2 4f w 5d l0 6p 6 7s 2 



For Group 2A metals, the same rule applies! Magnesium {Mg) is found in row 3 of the Periodic Table, and 
its valence electrons are found in the n = 3 energy level. Similarly, Radium {Ra) is found in row 7 of the 
Periodic Table, and its valence electrons are found in the n = 7 energy level. 

So far so good - but does the same pattern apply to the Group 3A - 8A elements (also known as Groups 
13-18). Let's find out by writing the electron configuration for the Group 3A element in each row. 



row 
row 
row 
row 
row 



2 
3 
4 
5 
6 



B: 
Al: 
Ga 
In : 
Tl: 



ls 2 2s 2 2p 1 

ls 2 2s 2 2p 6 3s 2 3p l 

ls 2 2s 2 2p Q 3s 2 3p^As 2 3d w Ap l 

ls 2 2s 2 2p 6 3s 2 3p fi 4s 2 3d 10 4p 6 5s 2 4d 10 5p 1 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d w 5p 6 6s 2 5d w 6p 1 



Even though the valence electrons in Group 3A elements are found in both s and p orbitals, it turns out 
that an element's row still corresponds to the energy level of that element's valence electrons. For example, 
Gallium {Ga) is found in row 4 of the Periodic Table, and its valence electrons are found in the n = 4 
energy level. Likewise, Thallium {Tl) is found in row 6 of the Periodic Table, and its valence electrons are 
found in the n = 6 energy level. 

It really does seem as if we can predict the energy level of an element's valence electrons using the row 
number for that element in the Periodic Table. Let's try one last example, though, just to be sure by 
writing the electron configuration for the Group 7A element in each row. 



265 



www.ckl2.org 



row 
row 
row 
row 
row 



B: 
Al: 
Ga : 
In : 
Tl: 



ls 2 2s 2 2p 5 

ls 2 2s 2 2p 6 3s 2 3p 5 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 5 

ls 2 2s 2 2p Q 3s 2 3p^s 2 3d w Ap^s 2 Ad l %p'' 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 5p 6 6 S 2 5d 10 6p 5 



Once again, an element's row can be used to determine the energy level of that element's valence electrons. 
Chlorine (CI), for instance, is found in row 3 of the Periodic Table, and its valence electrons are found 
in the n = 3 energy level. Similarly, Iodine (/) is found in row 5 of the Periodic Table, and its valence 
electrons are found in the n = 5 energy level. 

You should make note of one final point when it comes to energy levels and how they relate to the Periodic 
Table. Our rule for determining the energy level of an element's valence electrons using the element's row 
in the Periodic Table works for Group 1A - 8A elements. This rule doesn't apply, however, to the Group 
IB - 8B (also known as Groups 3 - 12) elements. The elements in that lower portion of the Periodic Table 
(the middle portion of the Periodic Table 'blacked out' in Figure 1) are known as transition metals. They 
behave differently, and you can't apply the same rules to them as far as valence electrons are concerned. 
The same goes for the isolated lower portion of the Periodic Table (also 'blacked out' in the figure of the 
Periodic Table above). This block contains elements known as lanthanides and actinides. Like transition 
metals, lanthanides and actinides do not obey the same rules as the Group 1A - 8A elements when it comes 
to valence electrons and valence electron energy levels. 

Hydrogen and Helium Occupy the First Period 

You probably noticed that, in the last section, we didn't mention the first row at all. Instead, we always 
started from row 2. The first row in the Periodic Table is a 'special row' for several reasons. To begin with, 
the first row of the Periodic Table contains only two elements - hydrogen and helium. Can you figure out 
why there are only two elements in the first row? 

According to what we learned in the last section, an element's row number corresponds to the energy level 
of that element's valence electrons. Therefore, the first row must contain elements with valence electrons 
in the n = 1 energy level. By now you should know that there is only one orbital in the n = 1 energy level. 
That orbital, of course, is the Is orbital. Hydrogen has one valence electron in the Is orbital, (its electron 
configuration is Is 1 ), and helium has two valence electrons in the Is orbital (its electron configuration is 
Is 2 ). Since it's impossible to fit more than two electrons into the Is orbital, atoms with a total of three or 
more electrons must have valence electrons in an energy level with n = 2 or greater. Clearly, then, atoms 
with a total of three or more electrons do not belong in the first row of the Periodic Table. 

The first row of the Periodic Table is also special because its elements have special properties. Hydrogen, 
for example, is not a metal like the rest of the Group 1A elements. Instead, hydrogen atoms react with each 
other and form what's known as hydrogen gas, H<i- As was mentioned in the last lesson, some scientists 
will put hydrogen in a category all by itself, rather than including it at the top of the 1A column. We 
won't do that here, but you should always keep in mind the fact that hydrogen is 'different,' and that you 
shouldn't compare the hydrogen's chemical properties with the chemical properties of the other Group 1A 
elements. 

Helium is also a special atom. You might wonder why helium appears at the far right-hand side of the 
Periodic Table, rather than right next to hydrogen. Again, helium's placement in the Periodic Table reflects 
its special chemical properties. Earlier you learned that Group 8A elements were 'inert' and that includes 



www.ckl2.org 



266 



helium. Even though helium only has two valence electrons, while the rest of the Group 8A elements have 
eight valence electrons, helium is placed on the top of the 8A column since helium's chemical behavior is 
similar to the chemical behavior of the other Noble Gases because it has a completed outer energy level. 

The s Sublevel Block on PT 



So far we know that, with the exception of transition metals, lanthanides and actinides, we can use the 
row in which an element is found to determine the energy level of that element's valence electrons. Can 
the organization of the Periodic Table, and the placement of an element within the Periodic Table, tell us 
anything else about the elements electron configuration? The answer is - 'y es -' Remember that the highest 
energy valence electrons in Group 1A and Group 2A elements are always in s orbitals. In fact, the only 
valence electrons in Group 1A and Group 2A elements are in s orbitals! Lithium, (Li) for instance, has 
the electron configuration ls 2 2s 1 . Notice that lithium's single valence electron is in an s orbital. Similarly, 
magnesium (Mg) has the electron configuration ls 2 2s 2 2p^'&s 2 . Again, notice that magnesium's two valence 
electrons are in an s orbital. Since all of the valence electrons in Group 1A and Group 2A elements exist 
in s orbitals, the first two columns of the Periodic Table (columns 1A and 2A) are known as the 's 
sublevel block.' The s sublevel block is shown in the figure below. Notice that the s sublevel block consists 
of all of the metals from Li down to Fr in column 1A, and all of the metals from Be down to Ra in column 
2A. Hydrogen is not included in the s sublevel block, again, because of its special properties. 

As pictured below, the s sublevel block of the Periodic Table includes the Group 1A metals and the Group 
2A metals. (Source: CK-12 Foundation. CC-BY-SA) 



JL 
















He 


JJ. 


Be 

BHM.JUM 


S Block 
/ 


JL 


',.?,_, 


'_H 





'f 


Nc 


Na 


Mg 


"a; 

1LUMIKUM 


"Si 


"f 


JL 


"ci 

CH1UHFNL 


Ar 


K 


Ca 


Sc 


"li 


"v 


"Cr 

CHROMIUM 


Mn 

UlNGlNir.C 


Fe 


Co 


"nj 


Cu 


Zn 


Ga 


Go 


As 


"Se 


Br 


Kr 


Rb 


Sr 


Y 


"zr 


Nb 


Mo 


"Tc 


4 Ru 


Rh 


"Pd 


k i 


"Cd 


In 


Sn 


"Sb 


T e 


I 


Xe 


Cs 


Ba 

G1F1IUM 


La-Lu 


Hf 


la 


W 


Re 


Os 


Ir 


Pt 


Au 


I 


Tl 


Pb 


Bi 


Po 


At 


Rn 


Fr 

FIUUCIUM 


Ra 


Ac-Lr 


If 


Db 


if. 


Bh 


Hs 


Nit 


Ds 


% 


Cn 


Gut 


Unci 


Uup 


Uuh 


Uus 


Uuo 



La 



Ac 



Ce 



Th 



Pr 



Pa 



Nd 



Pm 



Np 



Sm 



Pu 



Eu 



Am 



Gd 



Cm 



Tb 



Bk 



Cf 



Ho 



Es 



Er 



Fm 



Tm 



Md 



Yb 



No 



Lu 



Lr 



The p Sublevel Block on PT 

What can we say about the valence electrons in Group 3A- Group 8A elements? In particular, what can 
we say about the highest energy valence electrons? If you look carefully, you'll notice that for Group 3A 
Group 8A (or if you prefer, Groups 13-18) elements, the highest energy valence electrons are always in p 
orbitals. Boron, (B) for instance, has the electron configuration ls 2 2s 2 2p 1 . While boron has both 2s and 
2p valence electrons, the 2p valence electrons are higher in energy. Similarly, Krypton (Kr) has the electron 
configuration ls 2 2s 2 2p^?>s 2 ?>p & As 2 ?>d w Ap & . Again, while krypton has both 4s and \p valence electrons, the 
4p valence electrons are higher in energy. Since the highest energy valence electrons in Group 3A Group 
8A elements exist in p orbitals, the final six columns of the Periodic Table (columns 3A through 8A) 
are known as the 'p sublevel block.' The p sublevel block is shown in the figure below. Additionally, 



267 



www.ckl2.org 



as illustrated in the figure below the p sublevel block consists of all of the elements from B down to 77 in 
column 3A, all of the elements from C down to Pb in column 4A, all of the elements from TV down to Bi 
in column 5A, all of the elements from O down to Po in column 6A, all of the elements from F down to At 
in column 7A and, finally, all of the elements from Ne down to Rn in column 8A. Helium is not included 
in the p sublevel block, which should make sense, since helium has no p electrons! 

As illustrated below, the p sublevel block of the Periodic Table includes the Group 3A - Group 8A elements. 
(Source: CK-12 Foundation. CC-BY-SA) 



JL 
















He 


'Li 


Be 


P Block 
\ 


B 


JL. 


JL 


jo 


F 


Ne 


Na 


% 


Al 


Si 


"p 

FMOiPHUBUi 


"s 


PL 

CNLUBINE 


Ar 


"k 


Ca 


Sc 


I' 

TITtMUM 


"v 


CHBOMIUM 


Mn 

M1NG1KESE 


Fe 


Co 


"nj 


Cu 


Zn 


Ga 


Go 


As 


Se 


Br 


Kr 


Rb 


"Sr 


Y 


"Zr 


Nb 


Mo 


Tc 


Ru 


Rh 


Pd 


\i 


Cd 


In 


Sn 


"Sb 


To 


",!,. 


Xe 


Cs 


B| 


La-Lu 


Hf 


"la 


"w 


Re 


Os 


"lr 


"ft 


Au 


Hg 


"ti 


"Pb 


Bi 


Po 


"aj 


Rn 


"Vr 


Ra 


Ac-Lr 


Rf 


Db 


% 


lh 


Hs 


Mt 


Ds 


'l 


Cn 


Uut 


Unci 


Uup 


Uuh 


Uus 


Uuo 



LANTHANIDES 


La 


Ca 


"Pr 


Nd 


Pm 


Sm 

SIHWIUM 


Eu 


Gd 


T b 


1. 


Ho 


Er 


Tm 


Yb 


"Lu 


ACTIHIDES 


Ac 


T h . 


Pa 


U 

UFUWUM 


1 


"Pu 


Am 


Cm 


Bk 


"Cf 


"Es 


Fm 


Md 


No 


lr 



Just as the Periodic Table has an s sublevel block and a p sublevel block, it also has a d sublevel block and 
an / sublevel block. Defining valence electrons in the d and / sublevel blocks can be more difficult but, in 
general, most of the high energy valence electrons in the d sublevel block are in d orbitals while most of 
the high energy valence electrons in the / sublevel block are in / orbitals. 

As illustrated in the figure below, the Periodic Table is divided into the s sublevel block, the p sublevel 
block, the d sublevel block, and the / sublevel block. (Source: CK-12 Foundation. CC-BY-SA) 



H 


PERIODIC TABLE OF ELEMENTS 


'He 












Li 


Be 


S B Io <* P Block 

/ \ 

D Block 

1 


JL 


Is.. 


JL 


'o 


£ 


Ne 


Na 


% 


"aj 

ULUHHUM 


Si 

SILICON 


"p 


"s 


"pi 


"Ar 


"k 


Ca 


Sc 


u 


"v 


"ft 


Mn 


"Fe 


Co 


"Nj 


"Cu 


Zn 


Ga 


be 


As 


Se 


"lr 


Kr 


"Rb 


"Sr 


Y 


Zr 


Nb 


Mo 


*Tc 


Ru 


Rh 


Pd 


*! 


"Cd 


"in 


"Sn 


*Sb 


Te 


",!., 


54 

Xe 


"Cs 


Ba 


La-Lu 


Hf 


"la 


"w 


Re 


Os 


"lr 


Pt 


Au 


I 


JL 


"Pb 


Si 


Po 


"At 


Rn 


Fr 


"Ra 


Ac-Lr 


Rf 


Db 


H 


lh 


Hs 


Mt 


Ds 


J! 


Cn 


Uut 


Uuq 


Uup 


Uuh 


Uus 


Uuo 



LANTHANIDES 


la 


"Ce 


Pr 


Nd 


Pm 


Sm 


*Eu 

EURDPUIH 


*Gd 


Tb 


SL 


Ho 


'*Er 


Tm 


Yb 

YTTEBBIUM 


Lu 


ACTINIDES 


Ac 


T h 


Pa 


U 


1 


Pu 


Am 


Cm 


Bk 


"Cf 


"Es 


Fm 


Md 


No 


lr 



A complete summary of the Periodic Table, and the information it contains about an element's valence elec- 
trons is shown in the figure below. (Source: http://en.wikipedia.Org/wiki/File:PTable_structure. 
png. GNU-FDL) 



www.ckl2.org 



268 



MAIN GROUP ELEMENTS 




Earlier you learned that the Periodic Table was a convenient way to summarize all of the information that 
scientists know about the different elements found in our world. The Periodic Table probably even looked 
funny to you, because you had no way of understanding what its shape and organization meant. Now that 
we've discussed electron orbitals and electron energy levels, though, the Periodic Table shouldn't seem so 
strange anymore. In fact, the shape of the Periodic Table actually reflects the way in which electrons are 
organized in atoms of the different elements. 

The following webpage, Periodic Table (http://www.iun.edu/~cpanhd/periodictable.html) contains 
an interactive periodic table that allows you to click on the element and the web page then shows a number 
of physical constants for that element. The data includes, name, density, electron configuration, heat of 
vaporization, heat of fusion, melting point, boiling point, specific heat, and electronegativity. 

Lesson Summary 

• For Group 1A, 2A, 3A, 4A, 5A, 6A, 7A and 8A elements, the row in which an element is found 
corresponds to the energy level of that element's valence electrons. For example, lithium (Li) is in 
row 2, and its valence electrons are in the n = 2 energy level. 

• You can predict the energy level of an element's valence electrons using the element's row number in 
the Periodic Table. 

• Elements in the first row have special properties. 

• Hydrogen is not an alkali metal, and is usually found as a gas. 

• Helium is a Noble Gas, and exhibits chemical properties similar to the other Noble Gases found in 
Group 8A. 

• The Periodic Table can be divided into s,d,p and / sublevel blocks. 

• For elements in the s sublevel block, all valence electrons are found in s orbitals. 

• For elements in the p sublevel block, the highest energy valence electrons are found in p orbitals. 

Review Questions 

1 . Use the Periodic Table to determine the energy level of the valence electrons in each of the following 
elements. 

(a) B 



269 



www.ckl2.org 



(b) Ga 

(c) Rb 

(d) At 

(e) He 

2. Fill in the blanks: 

(a) B is in the level block of the Periodic Table 

(b) Sr is in the level block of the Periodic Table 

(c) Fe is in the level block of the Periodic Table 

(d) Cs is in the level block of the Periodic Table 

(e) O is in the level block of the Periodic Table 

3. Use the Periodic Table to determine the energy level and sublevel of the highest energy electrons in 
each of the following elements: 



(a) 


N 


(b) 


Ca 


(c) 


Rb 


(d) 


P 


(e) 


In 



4. Decide whether each of the following statements is true or false. 

(a) Li has valence electrons in the n = 1 energy level. 

(b) Si has valence electrons in the n = 3 energy level. 

(c) Ga has valence electrons in the n = 3 energy level. 

(d) Xe has valence electrons in the n = 5 energy level. 

(e) P has valence electrons in the n = 2 energy level. 

5. Match the element to the sublevel block it is found in: 

(a) C - i. s sublevel block 

(b) Cs - ii. p sublevel block 

(c) Ce - iii. d sublevel block 

(d) Cr - iv. / sublevel block 

6. The first row of the Periodic Table has: 

(a) 1 element 

(b) 2 elements 

(c) 3 elements 

(d) 4 elements 

(e) 5 elements 

7. Use the Periodic Table to determine which of the following elements has the highest energy valence 
electrons. 



(a) 


Sr 


(b) 


As 


(c) 


H 


(d) 


At 


(e) 


Na 



8. Use the Periodic Table to determine which of the following elements has the lowest energy valence 
electrons. 

(a) Ga 

(b) B 

www.ckl2.org 2T0 



(c) Cs 

(d) Bi 

(e) CI 

9. Which energy level does the first row in the d sublevel block correspond to? 

Further Reading / Supplemental Links 

• http : //learner . org/resources/series61 . html 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. The website has one 
video that applies to this lesson called The Periodic Table. 

• http : //dl . clackamas . cc . or . us/chl04-06/valence_electrons . htm 

• http : //chemed . chem . purdue . edu/genchem/topicreview/bp/ch8/index . php 

• http : //www . cem . msu . edu/~reusch/VirtTxt Jml/intro2 . htm 

• http : //en . wikipedia . org/wiki/Valence_electrons 

Vocabulary 

transition metals Elements in the d sublevel block (columns 1-B through 86) of the Periodic Table. 
The highest energy electrons in transition metals are found in d orbitals. 

lanthanides and actinides Elements in the / sublevel block of the Periodic Table. The highest energy 
electrons in lanthanides and actinides are found in / orbitals. 

s sublevel block The elements in the Periodic Table in columns 1A and 2A (excluding hydrogen). All 
valence electrons for elements in the s sublevel block are in s orbitals. 

p sublevel block The elements in the Periodic Table in columns 3A through 8A (excluding helium). 
The highest energy valence electrons for elements in the p sublevel block are in p orbitals. 

d sublevel block The elements in the Periodic Table in columns IB through 86 (also known as transition 
metals). 

/ sublevel block The elements in the lanthanide and actinide rows of the Periodic Table. 

noble gases Group 8A elements. These are elements found in the eight column of the Periodic Table. 
They are inert, which means that they are very non-reactive. 

inert Non-reactive. 



Image Sources 



271 www.ckl2.org 



Chapter 9 

Relationships Between the 
Elements 



9.1 Families on Periodic Table 

Lesson Objectives 



Describe the patterns that exist in the electron configurations for the main group elements. 
Identify the columns in the periodic table that contain 1) the alkali metals, 2) the alkaline earth 
metals, 3) the halogens, and 4) the noble gases, and describe the differences between each family's 
electron configuration. 

Given the outermost energy level electron configuration for an element, determine its family on the 
periodic table. 



Introduction 



With the introduction of electron configurations, we began to get a deeper understanding of the periodic 
table. An understanding of these electron configurations will prove to be invaluable as we look at bonding 
and chemical reactions. The orbital representation method for representing electron configuration is shown 
below. The orbital representation was learned in an earlier chapter but like many of the skills you learn in 
chemistry, it will be used a great deal in this chapter and in several chapters later in the course. 



www.ckl2.org 272 



gi 



7s 



6s 



/P 



Gp 



5p 



6d 



5d 



<W 



5f 



if 



4p 



3d 



3s 



2s 



3p 



-F 



In this lesson, we will focus on the connection between the electron configuration and the main group 
elements of the periodic table. We will need to remember the sub-level filling groups in the periodic table. 
Keep the following figure in mind. We will use it for the next two chapters. 

(Source: Richard Parsons. CC-BY-SA) 

Vertical Columns = "groups" or "families" 



Horizontal Rows = "Periods" 



1A 


2A 




3A 4A5AGA 


7A 


84 

















































































































































































J s block 

] p block 

□ d block 

I ] f block 



Alkali Metals Have One Electron in Their Outer Energy Level 

Elements Ending with s 1 = Alkali 

In the periodic table, the elements are arranged in order of increasing atomic number. In previous material 
we learned that the atomic number is the number of protons in the nucleus of an atom. For a neutral atom, 
the number of protons is equal to the number of electrons. Therefore, for neutral atoms, the periodic table 
is also arranged in order of increasing number of electrons. Take a look now at the first group or column 



273 



www.ckl2.org 



in the periodic table. It is the one marked "1A" in the Period Table figure below. The groups or families 
are the vertical rows of elements. The first group has seven elements representing the seven periods of the 
periodic table. Remember that a period in the periodic table is a horizontal row. Group 1A is the only 
group with seven elements in it. 

Group 1A of the Periodic Table (Source: Richard Parsons. CC-BY-SA) 



1A 


2A 




3 A 4A5A6A 


"A 


8A 



















































































































































































s block 

p block 

Lj d block 

□ f block 



Table 9.1: Electron Configurations for Group 1 Metals 



Element 



Atomic Number 



Electron Configuration 



Lithium (Li) 
Sodium (Na) 
Potassium (K) 
Rubidium (Rb) 
Cesium (Cs) 

Francium (Fr) 



3 
11 
19 
37 

55 

87 



ls 2 2s 2 2p 6 3s x 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 1 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 1 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d 10 

5p 6 6^ 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 

5p 6 6s 2 4f u 5d 10 6p 6 7s 1 



What do you notice about all of the elements in Group 1? They all have s 1 as the outermost energy level 
electron configuration. The whole number in front of the "s" tells you what period the element is in. For 
example sodium, Na, has the electron configuration ls 2 2s 2 2p 6 3s 1 , so it is in period 3. It is the first element 
of this period. 

This group of elements is called the alkali metals. They get their name from ancient Arabic (al kali) 
because "scientists" of the time found that the ashes of the vegetation they were burning contained a large 
amount of sodium and potassium. In Arabic, al kali means ashes. We know today that all alkali metals 
have electronic configurations ending in s 1 . You might want to note that while hydrogen is often placed in 
group 1, it is not considered an alkali metal. The reason for this will be discussed later. 

Alkaline Earth Elements Have Two Electrons in Their Outer En- 
ergy Level 

Elements Ending with s 2 — Alkaline Earth 

Taking a look at Group 2 A in Table 9.2, we can use the same analysis we used with group 1 to see if 



www.ckl2.org 



274 



we can find a similar trend. It is the second vertical group in the periodic table and it contains only six 
elements. 

What do you notice about all of the elements in group 2A? They all have an outermost energy level electron 
configuration of s 2 . The whole number in front of the "s" tells you what period the element is in. For 
example, magnesium, Mg, has the electron configuration ls 2 2s 2 2p e 3s 2 , so it is in period 3 and is the second 
element in that period. Remember that the s sublevel may hold two electrons, so in Group 2A, the s orbital 
has been filled. 

Elements in this group are given the name alkaline earth metals. They get their name because early 
"scientists" found that all of the alkaline earths were found in the earth's crust. Alkaline earth metals, 
although not as reactive as the alkali metals, are still highly reactive. All alkaline earth metals have 
electron configurations ending in s 2 . 

Table 9.2: Electronic Configurations for Group 2A Metals 



Element 



Atomic Number 



Electron Configuration 



Beryllium (Be) 
Magnesium (Mg) 
Calcium (Ca) 
Strontium (Sr) 
Barium (Ba) 

Radium (Ra) 



4 

12 
20 
38 
56 



ls 2 2s 2 

ls 2 2s 2 2p^s 2 

ls 2 2s 2 2p 6 3s 2 3/? 6 4s 2 

ls 2 2s 2 2p & 2,s 2 2,p & As 2 ?,d w Ap & hs 2 

ls 2 2s 2 2p 6 3s 2 3p 6 As 2 3d w Ap & hs 2 Ad w 

hp%s 2 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4/? 6 5s 2 4d 10 

5p%s 2 4f u 5d 10 6p 6 7s 2 



Noble Gases Have 8 Electrons in Their Outer Energy Level 

Elements Ending with s 2 p® = Noble Gases 

(Source: Richard Parsons. CC-BY-SA) 



1A 


2A 




3A 4A 5A &A 


7A 


SA 



















































































































































































s block 
p block 

□ d block 

□ f block 



The first person to isolate a noble gas was Henry Cavendish, who isolated argon in the late 1700 s. The 
noble gases were actually considered inert gases until the 1960s when a compound was formed between 
xenon and fluorine which changed the way chemists viewed the "inert" gases. In the English language, 
inert means to be lifeless or motionless; in the chemical world, inert means does not react. Later, the name 
"noble gas" replaced "inert gas" for the name of Group 8A. 



275 



www.ckl2.org 



When we write the electron configurations for these elements, we see the same general trend that was 
observed with groups 1A and 2A; that is, similar electron configurations within the group. 



Table 9.3: Electron Configurations for Group 8A Gases 



Element 



Atomic Number 



Electron Configuration 



Helium (He) 
Neon (Ne) 
Argon (Ar) 
Krypton (Kr) 
Xenon (Xe) 

Radon (Rn) 



2 
10 

28 
36 
54 

86 



ls z 

ls 2 2s 2 2p 6 

ls 2 2s 2 2p 6 3s 2 3p 6 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 

5p e 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 

Bp 6 6s 2 4f u 5d 10 6p 6 



Aside from helium, He, all of the noble gases have outer energy level electron configurations that are the 
same,ns 2 np & , where n is the period number. So Argon, Ar, is in period 3, is a noble gas, and would therefore 
have an outer energy level electron configuration of 3s 2 3p 6 . Notice that both the s and p sublevels are 
filled. Helium has an electron configuration that might fit into Group 2A. However, the chemical reactivity 
of helium, because it has a full first energy level, is similar to that of the noble gases. 

Halogens Have 7 Electrons in Their Outer Energy Level 

Elements Ending with s 2 p 5 = Halogens 
(Source: Richard Parsons. CC-BY-SA) 



1A 


2A 




3A 4A 5A SA 


7A 


SA 



















































































































































































s block 
p block 

□ d block 

□ f block 



The halogens are an interesting group. Halogens are members of Group 7A, which is also referred to as 
17. It is the only group in the periodic table that contains all of the states of matter at room temperature. 
Fluorine, F2, is a gas, as is chlorine, Ch- Bromine, Br2, is a liquid and iodine, I2, and astatine, At2, are both 
solids. What else is neat about Group 7A is that it houses four (4) of the seven (7) diatomic compounds. 
Remember the diatomics are H2,N2,02,F2,Ch, Br?, and h- Notice that the latter four are Group 17 
elements. The word halogen comes from the Greek meaning salt forming. French chemists discovered 
that the majority of halogen ions will form salts when combined with metals. We all know some of these 
already: LiF,NaCl,KBr, and Nal. 



www.ckl2.org 



276 



Taking a look at Group 7A in the figure, we can find the same pattern of similar electron configurations 
as found with group 1A, 2A, and 8A. It is the 17" 1 group in the periodic table and it contains only five 
elements. 

Table 9.4: Electron Configurations for Group 7A Elements 



Element 



Atomic number 



Electronic configuration 



Fluorine (F) 
Chlorine (CI) 
Bromine (Br) 
Iodine (/) 

Astatine (At) 



9 

17 

35 

53 

85 



ls 2 2s 2 2p 5 

ls 2 2s 2 2p 6 3s 2 3p 5 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 5 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 

5p 5 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 

5p 6 6s 2 4f M 5d 10 6p 5 



What is the general trend for the elements in Group 7A? They all have, as the outermost energy level 
electron configuration, ns 2 np 5 , where n is the period number. You should also note that these elements 
are one group away from the noble gases (the ones that generally don't react!) and the outermost electron 
configuration of the halogens is one away from being filled. For example, chlorine (CI) has the electron 
configuration [Ne] 3s 2 3p 5 so it is in period 3, the seventh element in the main group elements. The 
main group elements, as you recall, are equivalent to the s + p blocks of the periodic table (or the teal and 
purple groups in the diagram above). 

The Oxygen Family Has 6 Electrons in the Outer Energy Level 

Elements Ending with s 2 p 4 = the Oxygen Family 

(Source: Richard Parsons. CC-BY-SA) 



1A 


2A 




3A 4A 5A &A 


"A 


SA 



















































































































































































s block 
p block 

□ d block 

□ f block 



Oxygen and the other elements in Group 6 A have a similar trend in their electron configurations. Oxygen 
is the only gas in the group; all others are in the solid state at room temperature. Oxygen was first named 
by Antoine Lavoisier in the late 1700 s but really the planet has had oxygen around since plants were first 
on the earth. 

Taking a look at Group 6A in the figure below, we find the same pattern in electron configurations that we 
found with the other groups. Oxygen and its family members are in the 16 f/l group in the periodic table. 



277 



www.ckl2.org 



In Group 16, there are, again, only five elements. 

Table 9.5: Electron Configurations for Group 6A Elements 



Element 



Atomic Number 



Electron Configuration 



Oxygen (O) 
Sulfur (S) 
Selenium (Se) 
Tellurium (Te) 

Polonium (Po) 



16 

34 
52 

84 



ls 2 2s 2 2p A 

ls 2 2s 2 2p^s 2 ?,p± 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p' i 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d w 

5/? 4 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 

5p 6 6s 2 4f 14 5d 10 6p' i 



When we examine the electron configurations of the Group 6A elements, we see that all of these ele- 
ments have the outer energy level electron configuration of ns 2 np A . We will see that this similar electron 
configuration gives all elements in the group similar properties for bonding. 

These elements are two groups away from the noble gases and the outermost electron configuration is two 
away from being filled. Sulfur, for example, has the electron configuration ls 2 2s 2 2/? 6 3s 2 3/? 4 so it is in 
period 3. Sulfur is the sixth element in the main group elements. We know it is the sixth element across 
the period of the main group elements because there are 6 electrons in the outermost energy level. 

The Nitrogen Family Has 5 Electrons in the Outer Energy Level 



Elements Ending with s 2 p 3 = the Nitrogen Family 

(Source: Richard Parsons. CC-BY-SA) 



1A 


2A 




3A 4A 5A &A 


"A 


SA 



















































































































































































s block 
p block 

□ d block 

□ f block 



Just as we saw with Group 16, Group 5A has a similar oddity in its group. Nitrogen is the only gas in 
the group with all other members in the solid state at room temperature. Nitrogen was first discovered 
by the Scottish chemist Rutherford in the late 1700' s. The air is mostly made of nitrogen. Nitrogen has 
properties that are different in some ways from its group members. As we will learn in later lessons, the 
electron configuration for nitrogen provides the ability to form very strong triple bonds. 

Nitrogen and its family members belong in the 15 group in the periodic table. In Group 15, there are 



www.ckl2.org 



278 



also only five elements. 

Table 9.6: Electronic Configurations for Group 5A Elements 



Element 



Atomic Number 



Electron Configuration 



Nitrogen (N) 
Phosphorus (P) 
Arsenic (As) 
Antimony (Sb) 

Bismuth (Bi) 



7 

15 
33 
51 

83 



ls 2 2s 2 2p 3 

ls 2 2s 2 2p & 3s 2 3p 3 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 3 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d 10 

5p 3 

ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 

5p 6 6s 2 4f 14 5d 10 6p 3 



What is the general trend for the elements in group 5A? They all have, as the electron configuration in 
the outermost energy level, ns 2 np 3 , where n is the period number. These elements are three groups away 
from the noble gases and the outermost energy level electron configuration is three away from having a 
completed outer energy level. In other words, the p sublevel in the Group 15 elements is half full. Arsenic, 
for example, has the electron configuration ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 3 so it is in period 4, the fifth element 
in the main group elements. We know it is the fifth element across the period of the main group elements 
because there are 5 electrons in the outermost energy level. 

Lesson Summary 

• Families in the periodic table are the vertical columns and are also referred to as groups. 

• Group 1A elements are the alkali metals and all have one electron in the outermost energy level 
because their electron configuration ends in s . 

• Group 2A elements are the alkaline earth metals and all have two electrons in the outermost energy 
level because their electron configuration ends in s 2 . 

• Group 5A elements all have five electrons in the outermost energy level because their electron con- 
figuration ends in s 2 p 3 . 

• Group 6A elements all have six electrons in the outermost energy level because their electron config- 
uration ends in s 2 p 4 . 

• Group 7A elements are the halogens and all have seven electrons in the outermost energy level because 
their electron configuration ends in s 2 p 5 . 

• Group 8A elements are the noble gases and all have eight electrons in the outermost energy level 
because their electron configuration ends in s 2 p®. 

• Elements in group 8A have the most stable electron configuration in the outermost shell because the 
sublevels are completely filled with electrons. 

Review Questions 

1. If an element is said to have an outermost electronic configuration of ns 2 np 3 , it is in what group in 
the periodic table? 

(a) Group 3A 

(b) Group 4A 

(c) Group 5A 

(d) Group 7A 



279 



www.ckl2.org 



2. What is the general electronic configuration for the Group 8A elements? (Note: when we wish to indi- 
cate an electron configuration without specifying the exact energy level, we use the letter "»" to repre- 
sent any energy level number. That is, ns 2 np s represents any of the following; 2s 2 2p s , 3s 2 3p s , 4s 2 4/? 3 , 
and so on.) 

(a) ns 2 np 6 

(b) ns 2 np 5 

(c) ns 2 np 1 

(d) ns 2 

3. The group 2 elements are given what name? 

(a) alkali metals 

(b) alkaline earth metals 

(c) halogens 

(d) noble gases 



4. 



Using the diagram above, iden- 











B 




I 


T c 




















■ 


A 


E 












D 




H 


















G 

































tify: 

(a) The alkali metal by giving the letter that indicates where the element would be located and 
write the outermost electronic configuration. 

(b) The alkaline earth metal by giving the letter that indicates where the element would be located 
and write the outermost electronic configuration. 

(c) The noble gas by giving the letter that indicates where the element would be located and write 
the outermost electronic configuration. 

(d) The halogen by giving the letter that indicates where the element would be located and write 
the outermost electronic configuration. 

(e) The element with an outermost electronic configuration of s 2 p s by giving the letter that indicates 
where the element would be located. 

(f) The element with an outermost electronic configuration of s 2 p 1 by giving the letter that indicates 
where the element would be located. 

5. In the periodic table, name the element whose outermost electronic configuration is found below. 
Where possible, give the name of the group. 

(a) 5s 2 

(b) 4s 2 3d 10 4p 1 

(c) 3s 2 3p 3 



www.ckl2.org 



280 



(d) 5s 2 4d 10 5p 2 

(e) 3s l 

(f) l, 2 

(g) 6s 2 5d 10 6p 5 
(h) 4s 2 4p 4 

Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

group columns of the periodic table. 

period Horizontal rows of the periodic table. 

alkali metals Group 1 in the periodic table {Li,Na,K,Rb,Cs,Fr). 

alkaline earth metals Group 2 in the periodic table {Be,Mg,Ca,Sr,Ba,Ra). 

noble gases Group 18 in the periodic table (He,Ne,Ar,Kr,Xe,Rn). 

halogens Group 17 in the periodic table (F, CI, Br, I, At). 

main group elements Equivalent to the s + p blocks of the periodic table, also known as "representative 
elements." 

9.2 Electron Configurations 

Lesson Objectives 

• Convert from orbital representation diagrams to electron configuration codes. 

• Distinguish between outer energy level (valence) electrons and core electrons. 

• Use the shortcut method for writing electron configuration codes for atoms and ions. 

Introduction 

In the last section, we determined that the members of the chemical families (vertical columns on the 
periodic table) all have the same electron configuration in their outermost energy levels. In the next 
section, we will review the orbital representation for electron configuration and the electron configuration 
code. Then we will learn a shortcut for writing electron configuration codes. 

Shortcut for Writing Electron Configurations 

The diagram below represents the orbital representation diagram used in earlier chapters. The orbital 
representation diagram has a few different variations but all are used to draw electron configurations. Most 

281 www.ckl2.org 



show the orbitals in groups, as lines, boxes, or circles with each orbital having its own line (or circle) within 
each sublevel. 









o 9°oooooooooooo 

6s 

^ °Q° ooooo 

LJ 4d 

5s ooo 

o u - ooooo 

V_X 3d 

X ooo 

O 3p 

A ooo 

O 2p 

2s 

o 

Is 
The sublevel 2p has 3 orbitals (2p x ,2p y , and 2p z ) and each of these orbitals has its own line (or box). 



2p* 2 Pl 



2P 

Let's begin this section with the orbital box (or the orbital representation diagram) for a neutral atom. 
Draw the electronic configuration for potassium using the electron configuration diagram below. Remember 
that potassium is element number 19 so has 19 electrons. Every line in the energy diagram below holds 2 
electrons of opposite spins. 



www.ckl2.org 282 



a 

M 
fi 
(0 
A 
4) 
I* 
O 




ooo 

6p 

o 

6s 

°9° ooooo 

U> 4d 

5s OOO 

q u v u ooooo 
* ©®© 

O 3p 

A©©© 

© 2 P 



2s 

© 

Is 

Now we can simplify this by excluding all of the levels and sublevels that are not in use or do not hold 
electrons. 



&o 

u 
V 

c 
w 

bJO 

C 

« 



" ©©© 

© 3p 

"©©© 
© 2 p 



2s 



© 



Is 

Writing the electron configuration code for potassium, we would get: 

K : ls 2 2s 2 2p 6 3s 2 3p 6 4s 1 . 

Remember, the noble gases are the elements that are nearly non-reactive, partially due to the fact that the 
outermost energy level electron configuration was full (ns 2 np 6 ). Look at potassium on the periodic table 
and then find the preceding noble gas. Drag you finger counting backward from potassium to the nearest 
noble gas. What do you get? Argon (Ar) is correct. Now, let's draw the electron configuration for Argon 
and place the electron configuration for potassium beside it. 

283 www.ckl2.org 



> 

EP 

-L 



ti 



T5 - 

It 

3s 



u 



H H H 



3p 



H H H 



Electron configuration for Argon 



tL 

EZ 



t 




4s 


H U H 


H 


3p 



3s 



t* 



t* t* t* 



Electron configuration for Potassium 



What do you notice looking at the two diagrams? Potassium has the same electron configuration as argon 
+ 1 electron. We can make this point when we write electronic configuration codes. 



2o r .2 „6Q 2o»,6 / |„l 



K : ls z 2s z 2p°3s z 3p°4s 



„2o„2 



2o„6 



Ar : ls z 2s z 2p°3s z 3p 



Therefore, 

K : [Ar]4j 1 . {This is known as the noble gas configuration code system or the "short-cut"} 

Let's try another. Draw and write the electron configuration of gallium (Go). Gallium is number 31 and 
therefore has 31 electrons. It is also in the 4 period, the third element in the period of main group 
elements. From this information, we know that the outermost electrons will be As 2 Ap l . 




3s 3p 



It has 3 valence electrons so it 
is family 3A (column 13). 



r 

In period 3 

Electron configuration of Gallium (Ga). 

Writing the electron configuration code for gallium, we get: Ga : ls 2 2s 2 2p e 3s 2 3p e 4:S 2 3d l0 4:p l . In numerical 
order, the electronic configuration for gallium is: Ga : ls 2 2s 2 2p e 3s 2 3p e 3d 10 4s 2 4:p 1 . 



www.ckl2.org 



284 



t 



<1. 


n 


H 


^ u r+ n n 

3d 




4s 


ti U H 




si 


u 


3p 




- 


3s 


U ti U 






H 

2s 


2P 





Is 

Electron configuration for Gallium [Ga; 

Now, go back to the previous noble gas. Again, we find the previous noble gas is argon. Remember argon 
has electronic configuration of ls 2 2s 2 2p e 3s 2 3p e . 



Ga : ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 1 



Ar : ls 2 2s 2 2p 6 3s 2 3p (: 



Therefore, 



Ga: [Ar]4s 2 3d 10 V. 



This seems to work well for neutral atoms, what about ions? Remember that a cation is formed when 
electrons are removed from a neutral atom and anions are formed when electrons are added to a neutral 
atom. But where do the electrons get added to or removed from? The electrons that are involved in 
chemical reactions are the ones that are in the outer shell. Therefore, if electrons are to be removed or 
added, they get removed from or added to the shell with the highest value of n. 

Let's now look at an example. Draw the electron configuration for Fe. Then, write the reaction for the 
formation of Fe 2+ . 



285 



www.ckl2.org 



H 

4s 

H 

Els 


H H H 

3p 


H 


H H H 

2p 



UJ_i_i_i 



3d 



H 



Electron configuration for Iron (Fe) 

The reaction for the formation of the Fe 2+ ion is: Fe — > Fe 2+ + 2e~. Therefore we have to remove 2 
electrons. From where do we remove the two electrons? The rule is to remove from the highest n value 
first which, in this case, means we remove from the electrons from the 4 s sublevel first. 



Fe -» Fe 2+ + 2e 



„2o„2r 



„2qj6 



2 0o 2r, 6q„2o 6q j6 



ls'2s'2p 3s'3p°4s'3cF -» ls'2s'2p 3s'3p°3d° + 2e 



Remember that when we write electron configuration codes using noble gas configurations we use the 
previous noble gas. In writing the electron configuration code for Fe or Fe 2+ , we would use the symbol 
[Ar] to substitute for ls 2 2s 2 2p 6 3s 2 3p 6 . 



Fe -» Fe 2+ + 2e~ 
[Ar]3d 6 4s 2 -» [Ar]3J 6 + 2e~ 



In addition to forming Fe 2+ ions, iron atoms are also capable of forming Fe 3+ ions. Extra stability is gained 
by electron configurations that have exactly half-filled d orbitals. The third electron that is removed from 
an iron atom to form the Fe s+ ion is removed from the d orbital giving this iron ion a d orbital that is 
exactly half-filled. The extra stability of this configuration explains why iron can form two different ions. 



Fe -» Fe 3+ + 3e 



[Ai]3ds 6 4:s 2 -» [Ar]3ds 5 + 3e~ 



Now what about negative ions, the anions? Remember that negative ions are formed by the addition of 
electrons to the electron cloud. Look at bromine, Br. Neutral bromine has atomic number 35 and therefore, 
holds 35 electrons. 



www.ckl2.org 



286 



n n t 



L f 



w u n n w 



& iii 
u 

Jilt it 

n. 

2s 

it 
ts 

Electron configuration for Bromine (Br) 

The reaction for the formation of the Br~ ion is: Br + e~ —* Br~. The bromide ion is formed by the addition 
of one electron. If you look at the electron configuration for bromine, you will notice that the 4p sublevel 
needs one more electron to fill the sublevel and if this sublevel is filled, the electron configuration for Br is 
the same as that of the noble gas krypton (Kr). When adding electrons to form negative ions, the electrons 
are added to the lowest unfilled n value. 



Br + e 
ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 5 + e 



Br~ 

ls 2 2s 2 2/3s 2 3/4s 2 3d 10 4/7 e 



OR 

Br + e~ -» Br 
ls 2 2s 2 2p^s 2 ?,p (i As 2 M w Ap h + e~ -» [Kr] 

Notice how Br~ has the same electron configuration as Kr. When this happens, the two are said to be 
isoelectronic. 

Sample Question: Write the electron configuration codes for the following atoms or ions using the noble 
gas electron configuration shortcut. 

(a) Se 

(b) Sr 

(c) O 2 

(d) Ca 2+ 

(e) Al 3+ 
Solution: 

(a) Se[Ar] : As 2 3d w Ap A 

(b) Sr[Kr] : 5s 2 

(c) O 2 - : [Ne] 

(d) Ca 2+ : [Ar] 



287 



www.ckl2.org 



(e) AP+ : [Ne] 



Core Electrons 



1A 

1 2A 
2 



3A 4A 5A 6A 7 a 
13 14 15 16 17 



8A 
18 



3 4 5 6 7 8 9 10 11 12 




When we look at the electron configuration for any element we can distinguish that there are two different 
types of electrons. There are electrons that work to do the reacting and there are those that don't. The 
outer shell electrons are the valence electrons and these are the electrons that are responsible for the 
reactions that the atoms undergo. The number of valence electrons increases as you move across the 
periodic table for the main group elements. 

In older periodic tables, the groups were numbered using an A and B group numbering system. The 
representative groups (s and p blocks) were numbered with A and the transition elements were numbered 
with B. The table on the right has both numbering systems. The A numbering is shown for you in the 
diagram in red; this is all we will be concerned with for this lesson. The number of valence electrons in the 
A groups increases as the group number increases. In fact, the number of the A groups is the number of 
valence electrons. Therefore, if an element is in group 3A, it has 3 valence electrons, if it is in group 5A, 
it has 5 valence electrons. 

All electrons other than valence electrons are called core electrons. These electrons are in inner energy 
levels. It would take a very large amount of energy to pull an electron away from these full shells, hence, 
they do not normally take part in any reactions. Look at the electron configuration for gallium (Go). It 
has 3 valence electrons and 28 core electrons. 



www.ckl2.org 



288 



t 



<1. 


n 


H 


^ u r+ n n 

3d 




4s 


ti U H 




si 


u 


3p 




- 


3s 


U ti U 






H 

2s 


2P 





Is 

Electron configuration for Gallium [Ga; 

Sample Question: How many core electrons and valence electrons are in each of the following? 

(a) V 

(b) As 

(c) Po 

(d) Xe 
Solution: 

(a) 23 y '■ Core Electrons = 18, Valence electrons = 5 
(b)33 As : Core Electrons = 28, Valence electrons = 5 

(c) 84 Po : Core Electrons = 78, Valence electrons = 6 

(d) 54 Xe : Core Electrons = 46, Valence electrons = 8 



Lesson Summary 

• An orbital box diagram can be used to draw electron configurations for elements. 

• The noble gas configuration system is the shorthand method for representing the electron configura- 
tion of an element. 

• The noble gas symbol represents all of, or most of, the core electrons. 

• For cations, the electrons are removed from the outermost sublevel having the greatest n value. 

• For anions, the electrons are added to the unfilled energy level having the lowest n value. 

• Isoelectronic species have the same electron configurations. 

• Using an orbital box diagram, the core electrons and the valence electrons can easily be visualized. 

• The valence electrons are those electrons that are in the outermost shell. 



Review Questions 

1. What is the difference between the standard electron configuration code and the electron configuration 
code using noble gas configurations? 



289 



www.ckl2.org 



2. The standard electron configuration is often called the ground state electron configuration. Why do 
you think this is so? 

3. How is it possible to have the same number of core electrons and valence electrons for two different 
ions? 

4. Draw the orbital representation for the electron configuration of potassium, K. 

5. Write the electron configuration code for potassium, K. 

6. Write the noble gas electron configuration code for potassium, K. 

7. How many core electrons and valence electrons are in potassium, Kl 

8. Write the electron configuration for potassium, K + . 

9. Write the noble gas electron configuration code for potassium, K + . 

10. With what species is K + isoelectronic? 

11. The electron configuration below is for which element^ 2 



(a) 


Sb 




(b) 


Sn 




(c) 


Te 




(d) 


Pb 




. The electron configuration below is 


for whi^ e eLe i njg 1 t 2 ? 6/? 4 


(a) 


Pb 




(b) 


Bi 




(c) 


Po 




(d) 


TI 




. What is the noble gas electron configuration for the bromine element? 


(a) 


ls 2 2s 2 2p 6 3s 2 3p 6 4:S 2 4:p 5 




(b) 


ls 2 2s 2 2p 6 3s 2 3p 6 3d l0 4s 2 4p 5 




(c) 


[Ai]4s 2 4p 5 




(d) 


[Ai]3d 10 4s 2 4p 5 




. Write the noble gas electron configuration for each of the following. 


(a) 


AI 




(b) 


N 3 - 




(c) 


Sr 2 + 




(d) 


Sn 2+ 




(e) 


I 




. How 


many core electrons and valence electrons are in each of the following? 


(a) Mg 




(b) 


C 




(c) 


s 




(d) 


Kr 




(e) 


Fe 





Vocabulary 

orbital box diagram A diagram for drawing the electron configurations where sub-levels are shown in 
groups (or even in boxes) and each orbital has its own line (or box) within each sub-level. 

isoelectronic Having the same electron configuration. 

www.ckl2.org 290 



core electrons Electrons that occupy energy levels below the outermost energy level. 
valence electrons Electrons that occupy the outer shell of the atom or ion. 

9.3 Lewis Electron Dot Diagrams 

Lesson Objectives 

• Explain the meaning of an electron dot diagram. 

• Draw electron dot diagrams for given elements. 

• Describe the patterns of electron dot diagrams in the periodic table. 

Introduction 

This chapter will explore yet another shorthand method of representing the valence electrons. The method 
explored in this lesson will be a visual representation of the valence electrons. We will, as we observed 
in the previous lesson, finish the lesson with a look at how this visual representation flows in a pattern 
throughout the periodic table. 

A Simplified Way to Show Valence Electrons 

As defined earlier in this chapter, core electrons are all the electrons except the valence electrons and valence 
electrons are the electrons in the outermost energy level. Valence electrons are the electrons responsible 
for chemical reactions. Here is the electron configuration for sodium. 



- 

i 
■11 

b 



t 



3s- 



u u u 



"sr 



U 



Electron configuration for Sodium (Na) 



The electron configuration is Na : ls 2 2s 2 2p e 3s 1 . The core electrons are ls 2 2s 2 2p®. The valence electron 
is 3s 1 . One way to represent this valence electron, visually, was developed by Gilbert N. Lewis. These 
visual representations were given the name Lewis electron dot diagrams. Lewis suggested that since 
the valence electrons are responsible for chemical reactions and the core electrons are not involved, we 
should use a diagram that shows just the valence electrons for an atom. 



291 



www.ckl2.org 



To draw a Lewis electron dot diagram for sodium you can picture the symbol for sodium in a box with 
the box having four sides. Each side of the box represents either the s or one of the three p orbitals in 
the outermost energy level of the atom. The first and second valence electrons are placed on the side 
representing the s orbital and the next electrons are placed in the p orbitals. Just as the electrons are 
placed singly before being doubled up in the orbital representation, so they are placed one at a time in the 
p orbitals of an electron dot formula. A single dot represents one valence electron. Thus, the Lewis dot 
formula for sodium is: 

Na* 

Look at the electron configuration for magnesium. Magnesium is element #12. 



b£i 

C 

'« 

b 



u 

3b 



U H U 



H 
H 

Is 

Electron configuration for Magnesium (Mg) 



To draw the Lewis electron dot diagram we picture in our minds the symbol for Mg in a box with all of its 
core electrons (i.e., ls 2 2s 2 2p (i ). Then we place the valence electrons around the sides of the box with each 
side representing an orbital in the outermost energy level. How many valence electrons does magnesium 
have? Correct, there are 2 valence electrons, 3s 2 . 

Therefore the Lewis electron dot formula for magnesium is: 

M g : 

Look at the electron configuration for chlorine. 



www.ckl2.org 



292 



H H t 



H 

3s 


3p 


H 


2p 



ffl 

Of 

b 

H 

Electron configuration for Chlorine (CI) 

The electron configuration for chlorine could be written as: CI : ls 2 2s 2 2p®3s 2 3p 5 . The core electrons would 
be ls 2 2s 2 2p 6 while the valence electrons would be in the third shell (or where n = 3). Therefore, chlorine 
has 7 valence electrons. The Lewis electron dot diagram would look like the following: 

• • 

:ci* 

• • 

Notice that the electrons are in groups of two. Think of the chlorine in a box and the box has 4 sides. 
Each side can have 2 electrons on it. Therefore there can be a maximum of 2 x 4 = 8 electrons normally 
on any Lewis electron dot diagram. 

Sample question: Write the Lewis electron dot formula for: 

(a) Oxygen 

(b) Sulfur 

(c) Potassium 

(d) Carbon 

Solution: 

(a) Oxygen has the electron configuration: ls 2 2s 2 2p 4 , therefore there are 2 core electrons and 6 valence 
electrons. The Lewis electron dot formula is: 



:o* 



(b) Sulfur has the electron configuration: ls 2 2s 2 2p®3s 2 3p 4 , therefore there are 10 core electrons and 6 
valence electrons. The Lewis electron dot formula is: 



293 



www.ckl2.org 



:s * 

(c) Potassium has the electron configuration: 1j 2 2j 2 2/? 3j 2 3/? 6 4j , therefore there are 18 core electrons and 
1 valence electron. The Lewis electron dot formula is: 

K* 

(d) Carbon has the electron configuration: ls 2 2s 2 2p 2 , therefore there are 2 core electrons and 4 valence 
electrons. The Lewis electron dot formula is: 



All the Elements in a Column Have the Same Electron Dot Dia- 
gram 

In the previous lesson, it was shown that using the earlier version for numbering the periodic table, we 
could see a pattern for finding the number of valence electrons in each of the groups in the main group 
elements. Since the Lewis electron dot diagrams are based on the number of valence electrons, it would 
hold true that the elements in the same group would have the same electron dot diagram. In other words, 
if every element in Group 1A has 1 valence electron, then every Lewis electron dot diagram would have 
one single dot in their Lewis electron dot diagram. Take a look at the table below. 

Table 9.7: Lewis Electron Dot Diagrams for Group 1A Elements 

Element # Valence e~ Lewis electron dot diagram 

Hydrogen (H) 1 

Lithium (Li) 1 



Sodium (Na) 

Potassium (K) 
Rubidium (Rb) 

Cesium (Cs) 
Francium (Fr) 



H* 
Li- 

Na 

K* 

Rb* 

Cs* 

Fr- 



www.ckl2.org 294 



The same pattern holds for all elements in the main group. Look at Table 9.8 for the Lewis dot diagrams 
for Group 2 A elements. 

Table 9.8: Lewis Electron Dot Diagrams for Group 2A Elements 



Element 



# Valence e~ 



Lewis electron dot diagram 



Beryllium (Be) 
Magnesium (Mg) 
Calcium (Ca) 
Strontium (Sr) 
Barium (Ba) 
Radium (Ra) 



Be 
Mg 
Ca 
Sr 
Ba 
Ra 



A similar pattern exists for the Lewis electron dot diagrams for Group 3A. 

Table 9.9: 



Element 



# Valence e 



Lewis electron dot diagram 



Boron (B) 
Aluminum (Al) 
Gallium (Ga) 
Indium (In) 

Thallium (Tl) 



B* 

•• 
Al* 

•• 
Ga* 

•• 
In- 

• • 
Tl* 



Lesson Summary 

• A Lewis Electron Dot Diagram is used to visually represent only the valence electrons. 

295 www.ckl2.ors 



• The core electrons are symbolized using the symbol of the element with the valence electron dots 
surrounding the element symbol. 

• The maximum number of valence electron dots in the Lewis electron dot diagram is 8. Two electrons 
can go one each side (top, bottom, left, and right). 

• Electrons are added to the electron dot formula by placing the first two valence electrons in the s 
orbital, then one in each p orbital until each p orbital has one electron, and then doubling up the 
electrons in the p orbitals. 

• Each element in a group has the same number of valence electrons and therefore has the same Lewis 
electron dot diagram. (Same number of dots, different symbol.) 

Review Questions 

1. What is the advantage of the Lewis electron dot diagram? 

2. What is the maximum number of dots in a Lewis's Electron dot diagram? 

3. Draw the Lewis electron dot diagram for lithium. 

4. Draw the Lewis electron dot diagram for calcium. 

5. Draw the Lewis electron dot diagram for bromine. 

6. Draw the Lewis electron dot diagram for selenium. 

7. Write the Lewis electron dot diagrams for each of the following. What observations can you make 
based of these diagrams? 

(a) Oxygen 2- ion 

(b) Sulfur 2- ion 

(c) Antimony 

(d) Aluminum 

8. Write the trend for the Lewis electron dot diagrams for Group 6A (or Group 16) by filling in the 
table below. 

Table 9.10: 

Element # Valence e~ Lewis Electron Dot Diagram 

Oxygen (O) 
Sulfur (S) 
Selenium (Se) 
Tellurium (Te) 
Polonium (Po) 



Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

Lewis Electron Dot Diagram A shorthand visual representation of the valence electrons for an ele- 
ment. (Lewis electron dot diagram for sodium with one valence electron: 

Xa" 



www.ckl2.org 296 



9.4 Chemical Family Members Have Similar Prop- 
erties 

Lesson Objectives 

• Explain the role of the core electrons. 

• Explain the role of valence electrons in determining chemical properties. 

• Explain how the chemical reactivity trend in a chemical family is related to atomic size. 

Introduction 

Combining what we have learned from the previous lessons, we can now determine how the core electrons 
and the valence electrons relate to the properties of families in the periodic table. We have looked at the 
trend of the number of valence electrons within each group and now we will expand this to include a look 
at the trends that exist between the sublevel of valence electrons and the relative ability of an element to 
react in a chemical reaction. We will explore the possibility of predicting chemical reactivity based on the 
number of valence electrons and the sublevel to which they belong. Let's begin our lesson with a review of 
the core electrons and discuss their application in chemical reactions. Following this, we will have a similar 
discussion about valence electrons. 

Core Electrons Can be Ignored in Determining an Element's 
Chemistry 

As learned in an earlier lesson, the core electrons are the inner electrons found in the electron configuration 
for the element. These electrons fill up the inner sublevels and are thus not responsible for bonding and 
are not involved in chemical reactions. The core electrons also are not directly responsible for determining 
the properties of the elements. Remember that the core electrons represent all of the electrons except for 
the valence electrons. 

Look at the electron configuration for beryllium (Be). 

Be : ls 2 2s 2 

There are two core electrons (Is 2 ) and two valence electrons (2s 2 ). 

When we look at the electron configuration for Selenium (Se), we see the following. 

Se : ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 4 

There are 28 core electrons (ls 2 2s 2 2p e 3s 2 3p®3d 10 ) and six valence electrons (4s 2 4/? 4 ). Again, none of the 
core electrons from either of these elements, or any element for that matter, will participate in chemical 
reactions. 

Sample question: What are the core electrons in each of the following? 

(a) : ls 2 2s 2 2p 4 

(b) In: [Kr]5s 2 4d 10 5p l 
Solution: 

(a) O: core electrons = Is 2 , therefore there are two of them. 

297 www.ckl2.org 



(b) In: indium has three valence electrons and all the rest are core electrons. The core electrons include 
all the electrons in the energy levels below n = 5. (Ad 10 are core electrons since they are held in a lower 
energy level.) 

Valence Electrons Determine Chemical Properties 

The valence electrons, unlike the core electrons, are responsible for all of the chemical reactions that take 
place between elements. They determine the properties of the elements. We have already learned that 
all metallic elements in the main group elements are able to lose their valence electrons since the energy 
requirements to do so are relatively low. The non-metallic elements (other than noble gases) are able to 
gain electrons readily because of high electron affinities. For elements in Group 3, the atoms will have to 
lose three electrons to have electron configurations similar to the previous noble gas. For elements in group 
6, they will have to gain two electrons to have similar electron configurations as their nearest noble gas. 
Earlier we determined that all elements in the same group have the same Lewis Electron Dot diagram. 
What this means is that all elements in the same group have the same number of valence electrons and 
will react the same way to gain or lose electrons when participating in reactions. The number of valence 
electrons determines what types of chemical properties the elements will have. 

When you look at Table 9.11, you can see that the Group 1A metals all have the same number of 
valence electrons and also have the same appearance. These metals also react the same way under similar 
conditions. 

Table 9.11: Electron Configuration and General Appearance for Group 1A Elements 

Element Electron Configuration Appearance 

Li ls 2 2s 1 Silvery-white metal 

Na ls 2 2s 2 2p 6 3s 1 Soft, silvery-white metal 

K ls 2 2.s 2 2p 6 3.s 2 3p 6 4.s 1 Soft, silvery-white metal 

Rb ls 2 2.s 2 2/7 6 3.y 2 3p 6 4.y 2 3rf 10 4/? 6 5.s 1 Soft, silvery-white metal 

Cs ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d w Soft, silvery-white metal with 

5p 6 6i 1 gold sheen 

Fr ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d w ? too short lived 

5p 6 6s 2 4f 14 5d 10 6s 1 

We can see that all of the elements in the Group 1A metals all have one valence electron in their outer 
energy levels. This means that they can lose this one electron and become isoelectronic with a noble gas 
configuration. Thus, elements in Group 1A will readily lose this electron because it takes very low energy 
to remove this one outer electron. 

As the atomic number increases for the elements in the alkali metal family group, the valence electron is 
further away from the nucleus. The attraction between the valence electron and the nucleus decreases as 
the atomic size increases. The further the electrons are away from the nucleus, the less hold the nucleus 
has on the electrons. Therefore the more readily the electrons can be removed and the faster the reactions 
can take place. So, the valence electrons will determine what reactions will occur and how fast they will 
occur based on the number of electrons that are in the valence energy level. All group 1A elements will 
lose one electron and form a 1+ cation. 

Looking at Table 9.12 for the electron configuration of the Group 2A metals, we see that the outer energy 
level holds two electrons for each of the metals in group 2. 



www.ckl2.org 298 



Table 9.12: Electron Configuration for Group 2 A Elements 



Element Electronic Configuration 

Beryllium (Be) ls 2 2s 2 

Magnesium (Mg) ls 2 2s 2 2p 6 3s 2 

Calcium (Ca) ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 

Strontium (Sr) ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 bs 2 

Barium (Ba) ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 5p 6 6s 2 

Radium (Ra) ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d 10 5p 6 6s 2 4f u 5d w 6p 6 7s 2 

As with the Group 1A metals, the group 2A metals will form cations by losing the s electrons, but this 
time they will lose two electrons. And, as with the Group 1A elements, the elements are more reactive as 
the atomic number increases because the s electrons are held further away from the nucleus. 

Consider the electron configurations for the elements in family 7A (the halogens). 

Table 9.13: Electron Configuration for Group 7A Elements 

Element Electronic Configuration 

Fluorine (F) ls 2 2s 2 2p 5 

Chlorine (CI) ls 2 2s 2 2p 6 3s 2 3p 5 

Bromine (Br) ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 5 

Iodine (/) ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d w 4p 6 5s 2 4d w 5p 5 



The elements in group 7A were placed in the same chemical family by Mendeleev because they all have 
similar chemistry, that is, they react with other substances in similar ways and their compounds with 
the same metals have similar properties. Now we know why they have similar chemical properties . . . 
because chemical properties are controlled by the valence electrons and these elements all have the same 
configuration of valence electrons. 

The trend for chemical reactivity (speed of reaction) for the non-metal families is the reverse of the situation 
in the metal families. The metallic families lose electrons and the largest atoms lose their electrons most 
easily so the larger the atom, the faster it reacts. In the case of the non-metals, such as the halogens, the 
atoms react by taking on electrons. In these families, the smaller atoms have the largest electron affinity 
and therefore take on electrons more readily. Hence, in the non-metals, the smaller atoms will be more 
reactive. 



Lesson Summary 

• Core electrons are the inner electrons that are in filled orbitals and sublevels. 

• Core electrons are not responsible for chemical reactivity and do not participate in chemical reactions. 

• Valence electrons are the outermost electrons, are responsible for determining the properties, and are 
the electrons that participate in chemical reactions. 

• Because the members of each group in the main group elements has the same number of valence 
electrons, there are similar properties and similar trends in chemical reactions found in the group. 

• For metals, chemical reactivity tends to increase with increases in atomic size because the outermost 
electrons (or the valence electrons) are further away from the nucleus and therefore have less attraction 
for the nucleus. 

299 www.ckl2.org 



Review Questions 

1. What is the difference between valence electrons and core electrons? 

2. Why would the valence electrons be responsible for the chemical reactivity? 

3. Which of the following pairs of elements would have the greatest similarities in terms of chemical 
properties? 

(a) Ca and Ca 2+ 

(b) Ca and Mg 

(c) Ca and K 

(d) Ca and Ar 

4. What is the correct number of core electrons in the phosphorous atom? 

(a) 3 

(b) 5 

(c) 10 

(d) 15 

5. What is the correct number of valence electrons in the iodine atom? 

(a) 4 

(b) 5 

(c) 6 

(d) 7 

6. For the valence electrons in Group 6A, what conclusions can you draw about the trend in chemical 
reactivity? 

7. For the valence electrons in Group 7A, what conclusions can you draw about the trend in chemical 
activity? 

Further Reading / Supplemental Links 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 3-8 is on Lewis Dot Structures. 

• Lesson 1-8 is on Elemental Names and Symbols. 

• http : //www . f ordhamprep . org/gcurran/sho/sho/lessons/lesson31 . htm 

Vocabulary 

noble gas core When working with noble gas electronic configurations the core electrons are those 
housed in the noble gas symbolic notation. 

9.5 Transition Elements 

Lesson Objectives 

• Define transition metals. 

• Explain the relationship between transition metals and the d sublevels. 

• State the periods that contain transition metals. 

• Write electron configurations for some transition metals. 

www.ckl2.org 300 



Introduction 

The main group elements are actually in two groups: the s block and the p block. These two, let's call them 
sub groups, are separated in the periodic table by a special group of highly colorful compounds known as 
the transition metals. The transition metals represent groups 3 through 12 of the diagram below and are 
formed as electrons are filling in the d orbitals. This is why they are referred to as the d block. We will 
look at all of this in the lesson that follows. Let's begin!! 



Groups (or families) 



18 



>rk 


13 


14 15 


16 


17 




2 






d block 

3 4 5 6 7 8 9 10 11 12 














■o 3 
a 














oj 4 

Ph 






































5 




































6 




































7 




























1 


bl 


3Cl 




1 


■ 




■ 






1 


1 



Elements Whose Atoms are Filling d Sublevels 

The elements in groups 3 through 12 are called the transition metals. Sometimes this block of elements 
are referred to as the d block. They are called d block elements because the electrons being added in this 
block of elements are being added to the d orbitals. Look at the electron configurations for Scandium (Sc), 
Titanium (Ti), and Vanadium (V), the first three transition metals of the first row in the d block. 

215 c:ls 2 2s 2 2p (i 3s 2 3p <i 4s 2 3d 1 
22Ti:ls 2 2s 2 2p 6 3s 2 3p 6 As 2 3d 2 
23V:ls 2 2s 2 2p s 3s 2 3p s 4s 2 3d 3 

Notice that as the atomic number increases, the number of electrons in the d sublevels increases as well. 
This is why the phrase d block was started for the transition metals. The transition metals were given 
their name because they had a place between the 2A group (now Group 2) and the 3A (now Group 13) in 
the main group elements. Therefore, in order to get from calcium to gallium in the periodic table, you had 
to transition your way through the first row of the d block ( Sc — » Zn). To get from strontium to indium, 
you had to transition your way through the second row of the d block ( Y — » Cd) . 

One interesting point about the transition metals that is worth mentioning is that many of the compounds 
of these metals are highly colored and used in dyes and paint pigments. Also, metallic ions of transition 
elements are responsible for the lovely colors of many gems such as jade, turquoise and amethyst. 

Sample question: Write the electron configuration code for Fe. 

Solution: 

26Fe: 1 s 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 6 



301 



www.ckl2.org 



Transition Metals Occur in Periods 4 — 7 

As stated before, transition metals are also called the d block because they fill up the d sublevels. Therefore, 
the first three periods do not have transition elements because those energy levels do not have d sub- shells. 
The transition metals have atomic numbers greater than 20 because the first d electron appears in element 
#21 after the Is, 2s, 2p, 3s, 2>p, and 4s orbitals are filled. 



1 


1 


2 




Croups (or families) 


13 


14 15 


16 


17 


18 


2 






3 


(I block 

4 5 6 7 8 9 10 11 12 














13 3 

o 














<u 4 






































5 






































6 






































7 










































t OIOCK 









Lesson Summary 

• Transition metals are those from the d block or those from Groups 3 through 12. 

• Compounds of transition metals typically are those that are highly colored. 

• Transition metals are found in Periods 4 through 7. 

Further Reading / Supplemental Links 

• http : //learner . org/resources/series61 . html 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. The website has one 
video that apply to this lesson. It is called The Atom and deals with the internal structure of the atom. 

Review Questions 

1. Write the electron configuration for zirconium, Zr. 

2. How many valence electrons does zirconium, Zr, have? 

3. Write the noble gas electronic configuration for platinum, Pt. 

4. How many valence electrons does platinum, Pt, have? 

5. Why do the d block elements only start in the fourth period? 

6. What do copper, silver, and gold have in common as far as their electron configuration? 

7. Which of these is the electron configuration for nickel? 

(a) [Kr]3d 8 4s 2 

(b) [Kr]3J 10 

(c) [Ar]3d 8 4s 2 



www.ckl2.org 



302 



(d) [Ar]3d 10 

8. How many d electrons are there in the electronic configuration for ruthenium? 

(a) 

(b) 6 

(c) 7 

(d) 17 

9. Write the electron configuration for Iridium, Ir. 

10. What are the valence electrons for Iridium, Ir? 

11. Write the noble gas electron configuration for mercury, Hg. 

12. How many valence electrons does mercury, Hg, have? 

Vocabulary 

transition metal Groups 3 through 12 or the d block of the periodic table. 

Labs and Demonstrations for Electron Configuration 
Teacher's Pages for Paramagnetism Lab 

a————- 

After discussing the electron configurations that produce paramagnetism, the student will investigate the 
paramagnetism (or lack thereof) with several compounds. 

Lab Notes Capsules for this lab can be pre-loaded with the manganese salts. This will save time with a 
large class, and is safer. Mn02 can stain the skin and is a mild oxidizing agent. MnO~ A is a strong oxidizing 
agent and can cause burns. 

If you have trouble with the procedure involving the balance due to balance sensitivity, you may try placing 
the gel capsule onto a smooth surface and bring a strong magnet towards it. The capsule may exhibit an 
attraction, repulsion, or no reaction. The paramagnetism of the material may be gauged ordinally. 

Answers to Pre-Lab Questions 

1. Paramagnetism is a weak attraction of a substance for a magnetic field due to the presence of unpaired 
electrons. The unpaired electrons each exert a weak magnetic field due to spin, and the spins are 
additive when unpaired. 

2. Paramagnetism is a weak attraction for a magnetic field, while diamagnetism is a weak repulsion to 
a magnetic field due to opposing electron spins within an orbital. 

3. The presence of d sublevel in transition metals gives nine opportunities for the sublevel to have an 
unpaired electron, not accounting for d 4 or d 9 promotion. This is more than the p sublevel or the s 
sublevel. 

Lab - Paramagnetic Behavior of Manganese Compounds 

Background Information 

303 www.ckl2.org 



When we construct electron configurations for the atoms and ions in the periodic table, one of the cardinal 
rules that chemists observe is the pairing of electrons in orbitals whenever possible. Substances with 
paired electrons display diamagnetism, a phenomenon that causes that material to be slightly repelled by 
a magnetic field. Atoms with unpaired electrons, known as paramagnetic substances, are weakly attracted 
to magnetic fields. One interesting example of a molecule that displays paramagnetism is the oxygen 
diatomic molecule. While a Lewis dot structure can be constructed pairing all twelve valence electrons 
into an oxygen - oxygen double bond with two pairs of non-bonded electrons on each oxygen atom, the 
molecule displays paramagnetism when poured between the poles of a strong magnet, attracted to the 
magnetic field. 

Pre-Lab Questions 

1. Explain what is meant by the term paramagnetic. 

2. Explain how paramagnetism differs from diamagnetism. 

3. Why might paramagnetic transition metal compounds be more prevalent than those of main group 
elements? 

Purpose 

The purpose of this experiment is to demonstrate the effects of a magnetic field on a paramagnetic material. 
Apparatus and Materials 

• MnC>2 

• Mn 2 03 
. KMnO^ 

• Styrofoam cups 

• Magnets 

• Empty gel capsules 

Safety Issues 

Manganese oxide is a strong oxidant and can irritate tissues. Potassium permanganate is also a powerful 
oxidizing agent and can irritate skin The use of eye protection is strongly recommended. 

Procedure 

Place about 1.0 gram of each of the materials to be tested in empty gel capsules. The capsules are then 
placed on the base of an inverted Styrofoam cup. A second Styrofoam cup is then placed beneath the first 
cup. The entire assembly is then placed on a top loading balance. Obtain the mass of the cup and gel 
capsule assembly. Now bring the magnet to the cup/capsule assembly and reassess the mass. Note the 
decrease in mass as the magnet attracts the capsule contents. 

Table 9.14: Data 

With Magnet Without Magnet 

Mass of the 

MnOi /capsule/cup 

assembly 

Mass of the 

Mn-i O3 /capsule/cup 

assembly 

www.ckl2.org 304 



Table 9.14: (continued) 



With Magnet 



Without Magnet 



Mass of the 

ATMnC^/capsule/cup 

assembly 



Post-Lab Questions 

1 . Which of the samples displayed the largest mass difference? 

2. How many unpaired electrons does Manganese possess in MnO^^- 

In Mrc 2 03? 
In KMnO A l 

9.6 Lanthanides and Actinides 

Lesson Objectives 

• Define the lanthanides and actinides. 

• Place the lanthanides and actinides in the periodic table. 

• Explain the importance of both the lanthanides and actinides. 

• Write electron configurations for lanthanides and actinides. 

Introduction 

To complete our look at the periodic table there is one more group we have to consider. This group holds its 
own unique position in any pictorial representation of the periodic table. The two rows that are generally 
placed underneath the main periodic table are called the lanthanide series and the actinide series. These 
two rows are produced when electrons are being added to f orbitals. Therefore, this block of elements are 
referred to as the "f block". The lanthanides are also occasionally referred to the rare earth elements. 
The f block, as shown in the figure below, are the two rows in red. 



1 


1 


2 


Croups (or families) 


13 


14 15 16 


17 


18 


2 






d block 

3 4 5 6 7 8 9 10 11 12 














■S 3 

5 














<u 4 

2rf 






































5 






































6 






































7 







































f block 




305 



www.ckl2.org 



Lanthanides and Actinides are Elements Filling the f Sublevels 

There is one group that we have neglected to mention throughout this chapter. This group belongs to a 
special group found almost disjointed, if you like, from the periodic table. They are given the names the 
Lanthanide Series and the Actinide Series or the lanthanides and the actinides. The lanthanides are 
an important group of elements. Most of them are formed when uranium and plutonium undergo nuclear 
reactions. The elements of the lanthanide series are also known as the rare earth elements. Both the 
lanthanides and the actinides make up what are known as the inner transition series. The / block is given 
this name because if the f block were placed in its proper numerical position in the periodic table, it would 
be in the transition metals between groups 2 and 3. 

The lanthanide series includes elements from number 58 to 71, which is 14 elements. The / sub-level 
contains seven orbitals and each orbital will hold two electrons. Therefore, it is possible to place 14 
electrons in the 4/ sub-level. 

The lanthanide series fills the 4/ sublevel as you move from cerium (Ce) to lutetium (L«). The same holds 
for the actinide series that runs from atomic number 90 through to number 103, again 14 elements. Thus, 
as you move from thorium (Th) at element number 90, you begin to fill up the 5/ sublevel and continue 
to fill up the 5/ sublevel until you finish the actinide series at lawrencium (Lr). 



1 


































18 


1A 


2 


















13 


14 15 16 


17 


8A 


H 


He 


HYDROGEN 


2A 


















3A 


4A 5A 6A 


7A 


HELIUM 


3 

Li 


4 

Be 


5 

B 


6 

c 


7 

N 


8 




9 

F 


10 

Ne 


imam 


■nuin 


3 


4 


5 


6 


7 


8 9 10 


ii 


12 


BOHON 


CARBON 


NITROGEN 


OXYGEN 


FLUORINE 


NEON 


Na 


%g 

MAGNESIUM 


13 

Al 


14 

Si 


15 

P 


16 

s 


13 

CI 


18 

Ar 


SODIUM 


3B 


4B 


5B 


6B 


7B 


1 »B 1 


IB 


2B 


ALUMINUM 


SILICON 


PHOSPHORUS 


SULFUR 


CHLORINE 


ARGON 


19 

K 


20 

Ca 


21 

Sc 


22 

Ti 


23 

V 


24 

Cr 


25 

Mn 


26 

Fe 


27 

Co 


28 

Ni 


29 

Cu 


30 

Zn 


31 

Ga 


32 

Ge 


33 

As 


34 

Se 


35 

Br 


36 

Kr 


POTASSIUM 


CALCIUM 


SCANDIUM 


TITANIUM 


VANADIUM 


CHROMIUM 


MANGANESE 


IRON 


COBALT 


NICKEL 


COPPER 


ZINC 


GALLIUM 


GERMANIUM 


ARSENIC 


SELENIUM 


BROMIUM 


KRYPTON 


37 

Rb 


38 

Sr 


39 

Y 


40 

Zr 


41 

Nb 


42 

Mo 


43 

Tc 


44 

Ru 


45 

Rh 


46 

Pd 


47 

A g 


48 

Cd 


49 

In 


50 

Sn 


51 

Sb 


52 

Te 


53 

1 


54 

Xe 


RUBIDIUM 


STR0N7.UM 


YTTRIUM 


ZIRCONIUM 


NIOBIUM 


MOLYBDENUM 


TECHNETIUM 


RUTHENIUM 


RHODIUM 


PALLADIUM 


SILVER 


CADMIUM 


INDIUM 


TIN 


ANTIMONY 


TELLURIUM 


IODINE 


XENON 


55 

Cs 


56 

Ba 


57-71 

La-Lu 


72 

Hf 


73 

Ta 


74 

w 


75 

Re 


76 

Os 


77 

lr 


78 

Pt 


79 

Au 


80 

H g 


Bl 

TI 


82 

Pb 


83 

Bi 


84 

Po 


85 

At 


86 

Rn 


CESIUM 


BARIUM 


LANTHANIDES 


HAFNIUM 


TANTALUM 


TUNGSTEN 


RHENIUM 


OSMIUM 


IRIDIUM 


PLATIUM 


GOLD 


MERCURY 


THALUUM 


LEAD 


BISMUTH 


POLONIUM 


ASTATINE 


RADON 


87 

Fr 


88 

Ra 


89-103 

Ac-Lr 


104 

Rf 


105 

Db 


106 

sg 


107 

Bh 


108 

Hs 


109 

Mt 


110 

Ds 


111 

Rg 


112 

Cn 


113 

Uut 


114 

Uuq 


115 

Uup 


116 

Uuh 


117 

Uus 


118 

Uuo 


FRANC IUM 


RADIUM 




FIUTHERFOFIDUM 


DUBNIUM 


SEABORGIUM 


BOHRIUM 


HA5SIUM 


MEITNERIUM 


«..«, .,,». 




COPERNICIUM 


UNUNTRIUM 


mammm 


UNUNPENTIUU 


i „.,... 


UHUNSEPII.. 


UN.K0C1IUM 



LANTHANIDES 


57 

La 


58 

Ce 


59 

Pr 


60 

Nd 


61 

Pm 


62 

Sm 


63 

Eu 


64 

Gd 


65 

Tb 


66 

Dy 


67 

Ho 


68 

Er 


69 

Tm 


70 

Yb 


71 

Lu 




LANTHANUM 


CERIUM 


PHASE0OYM.UM 


NEODYMIUM 


PROMETHIUM 


SAMARIUM 


EUROPIUM 


GADOLINIUM 


TERBIUM 


DYSPROSIUM 


HOLMIUM 


ERBIUM 


THULIUM 


YTTERBIUM 


LUTETIUM 


ACTINIDES 


89 

Ac 


90 

Th 


91 

Pa 


92 

u 


93 

Np 


94 

Pu 


95 

Am 


96 

Cm 


93 

Bk 


98 

Cf 


99 

Es 


100 

Fm 


101 

Md 


102 

No 


103 

Lr 




ACTINIUM 


THORIUM 


PROTACTINIUM 


URANIUM 


NEPTUNIUM 


PLUTONIUM 


AMERICIUM 


CURIUM 


BERKELIUM 


CALIFORNIUM 


EINSTEINIUM 


FERMIUM 


MENOELEVIUM 


NOBELUJM 


immam 



Lanthanides and Actinides Vary in Electron Filling Order 

The lanthanides and the actinides make up the f block of the periodic table. The lanthanides are the 
elements produced as the 4/ sublevel is filled with electrons and the actinides are formed while filling the 
5/ sublevel. Generally speaking, the lanthanides have electron configurations that follow the Aufbau rule. 
There are some variations, however, in a few of the lanthanide elements. We will expand a tiny portion 
of the periodic table below to show what happens to some of the electron configurations in the lanthanide 
and actinide series. 



www.ckl2.org 



306 



19 


20 


21 


22 


23 


K 


Ca 


Sc 


Ti 


V 


39.098 


40.Q7B 


44.956 


47.867 


50.942 


POTASSIUM 


CALCIUM 


SCANDIUM 


TITANIUM 


VANADIUM 


37 


38 


39 


40 


41 


Rb 


Sr 


Y 


Zr 


Nb 


B5.46B 


37.62 


88.906 


91.224 


92.905 


RUBIDIUM 


STRONTIUM 


YTTRIUM 


ZIRCONIUM 


NIOBIUM 


55 


56 


57-71 


72 


73 


Cs 


Ba 


La-Lu 


\Hf 


Ta 


CESIUM 


137.327 
BARIUM 


LANTHANIDES 


HAFNIUM 


TANTALUM 


87 


88 


89-103 


l 104 \ 


105 


Fr 


Ra 


Ac-Lr 




Db 


223.020 

FRANCIUM 


226.0254 
RADIUM 


ACTINIDES 


Ve3.ii3 \ 

RUTHKfORDlim 


262.114 
DUBNIUM 



57 

La 

. LANTHANUM 


58 

Cc 

CERIUM 


59 

Pr 

PRASEODYMIUM 


60 

Nd 

NEODVMIUM 


61 

Pm 

PRQMETHIUM 


62 

Sm 

SAMARIUM 


63 

Ell 

EUROPIUM 


64 

Gd 

GADOLINIUM 


89 

Ac 

227.027 
i ACTINIUM 


90 

Th 

232.038 
THORIUM 


91 

Pa 

231.036 
PROTACTINIUM 


92 

u 

2 3 B. 029 
URANIUM 


93 

Np 

237.01a 
NEPTUNIUM 


94 

Pu 

244.06+ 
PLUTONIUM 


95 

Am 

243.061 
AMERICIUM 


96 

Cm 

247.070 
CURIUM 



Look, for example, at the electron configuration for cerium, the first element of the lanthanide series. 



Cerium, Ce, is element number 58. 



58Ce: ls 2 2i- 2 2p 6 3.s' 2 3p 6 4i 2 3d 10 4p 6 5i 2 4rf 10 5p 6 6i' 2 5rf 1 4/ 1 



or 

58Ce:[Xe]6s 2 5d 1 4f 1 

Now look at the electronic configuration for praseodymium, an element used in the making of aircraft 
engines but also in lighting for making movies. Praseodymium, Pr, is element number 59 and has the 
following electron configuration. 

59Pr: l.? 2 2.s 2 2p 6 3s 2 3p 6 4i 2 3d 10 4p 6 5.s 2 4rf 10 5p 6 6i 2 4/ 3 

or 

59Pr:[Xe]6i 2 4/ 3 

Notice the d electron is no longer a part of the electron configuration. There are three lanthanide metals 
that have properties similar to the d block. These are cerium, Ce, lutetium, Lu, and gadolinium, Gd. All 
of these metals contain a d electron in their electron configuration. The rest, like praseodymium, simply 
fill the 4/ sublevel as the atomic number increases. 

Unlike the lanthanide family members, most of the actinide series are radioactive. Most of the elements 
in the actinide series have the same properties as the d block. Members of the actinide series can lose 
multiple numbers of electrons to form a variety of different ions. Table 9.15 shows the noble gas electron 
configuration for the elements of the actinide series. 



307 



www.ckl2.org 



Table 9.15: Noble Gas Electron Configuration for the Actinide Series 



Element 



Electron Configuration 



Thorium (Th) 
Protactinium (Pa) 
Uranium (U) 
Neptunium (Np) 
Plutonium (Pu) 
Americium (Am) 
Curium (Cm) 
Berkelium (Bk) 
Californium (Cf) 
Einsteinium (Es) 
Fermium (Fm) 
Mendelevium (Md) 
Nobelium (No) 
Lawrencium (Lr) 



[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 
[Rn 



7s 2 M 2 

7s 2 5f 2 6d 1 

7s 2 hf<od 1 

7s 2 5/ 4 6d 1 

7s 2 5f 

7s 2 5f 

7s 2 5f 7 6d 1 

7s 2 5f 

7s 2 5f w 

7s 2 5f u 

7s 2 5f 12 

7s 2 5f 13 

7s 2 5f 14 

7s 2 5f M 6d 1 



Lesson Summary 



The lanthanide and actinide series make up the inner transition metals. 

The lanthanide series fill up the 4/ sublevel and the actinide series fill up the 5/ sublevel. 

The first, middle, and last member of the lanthanide series have properties of the f block and the d 

block. 

Many of the actinide series have properties of both the d block and the f block elements. 



Review Questions 



1. Why are the f block elements referred to by some as inner transition elements? 

2. What do europium and americium have in common as far as their electron configuration? 

3. What is the electron configuration for Berkelium? 

(a) [Xe]7s 2 5/ 9 

(b) [Xe^s^fM 1 

(c) [Rn]7s 2 5/ 9 

(d) [Rv\7s 2 5fQd l 

4. How many f electrons are there in the electron configuration for einsteinium? 

(a) 

(b) 11 

(c) 14 

(d) 25 

5. Write the electron configuration for Ytterbium, Yb. 

6. What are the valence electrons for Ytterbium, Yb? What periods and sublevels are they in? 

7. Write the noble gas electronic configuration for uranium, U. 

8. What are the valence electrons for uranium, U? What periods and sublevels are they in? 

9. Write the electron configurations for neptunium and then for plutonium. Now write an explanation 
for what seems to be happening. 



www.ckl2.org 



308 



Further Reading / Supplemental Links 

• http : //en . wikipedia . org/wiki/Group_number_of _lanthanides_and_act inides 

• http : //en . wikipedia . org/wiki/Lanthanide 

Vocabulary 

lanthanide The rare earth elements found in the first period of the / block. These elements fill up the 
4/ sublevel. 

actinide The elements found in the second period of the / block. These elements fill up the 5/ sublevel. 

Image Sources 



309 www.ckl2.org 



Chapter 10 



Trends on the Periodic Table 



10.1 Atomic Size 

Lesson Objectives 



Define atomic radius. 

State the boundary issue with atomic size. 

Describe measurement methods for atomic size. 

Define the shieiding effect. 

Describe the factors that determine the trend of atomic size. 

Describe the general trend in atomic size for groups and for periods. 

Describe the trend of atomic radii in the rows in the periodic table. 

Describe how the trend of atomic radii works for transition metals. 

Use the general trends to predict the relative sizes of atoms. 



Introduction 

In the periodic table, there are a number of physical properties that are not really "similar" as it was 
previously defined, but are more trend-like. This means is that as you move down a group or across a 
period, you will see a trend-like variation in the properties. There are three specific periodic trends that we 
will discuss. The first three lessons of this chapter are devoted to the trend in atomic size in the periodic 
table. Following this we will discuss ionization energy and electron affinity. Each of these trends can be 
understood in terms of the electron configuration of the atoms. 

The actual trends that are observed with atomic size have to do with three factors. These factors are: 



1. The number of protons in the nucleus (called the nuclear charge). 

2. The number of energy levels holding electrons (and the number of electrons in the outer energy level). 

3. The number of electrons held between the nucleus and its outermost electrons (called the shielding 
effect). 



Comparative Atomic Sizes. (Source: CK-12 Foundation. CC-BY-SA) 
www.ckl2.org 310 



Family 
1A 2A 3A 4A 5A 6A 7A 8A 



He 



Be /R ,C 7 N ,0 y V x Nc 



/EKJ ,,- " /<- /i" /" ,1 

Li) 



g 3 



CL 




Al ^ Q ^ ^ © 



K Ca Ga Ge As 



Sc Bt Kr 



5 Rb 



Sr In Sri Sb Te 1 



Xe 



6 Cs 



Ba Tl Pb Bi Po At Rn 



Atoms Have No Definite Boundary 

The region in space occupied by the electron cloud of an atom is often thought of as a probability distri- 
bution of the electrons and therefore, there is no well-defined "outer edge" of the electron cloud. Atomic 
size is defined in several different ways and these different definitions often produce some variations in the 
measurement of atomic sizes. 

Atomic radius of H2 






H 



H 



311 



www.cki2.0rg 



Because it is so difficult to measure atomic size from the nucleus to the outermost edge of the electron cloud, 
chemists use other approaches to get consistent measurements of atomic sizes. One way that chemists define 
atomic size is by using the atomic radius. The atomic radius is one-half the distance between the centers 
of a homonuclear diatomic molecule (a diatomic molecule means a molecule made of exactly two atoms 
and homonuclear means both atoms are the same element). The figure below represents a visualization of 
the atomic size definition. 

The image on the right is a visual representation of the atomic radius of a hydrogen atom. The measurement 
would be taken as one-half the distance between the nuclei of the hydrogen atoms in a diatomic hydrogen 
molecule. 

How do we measure the size of the atom? Ernest Rutherford is famous for his experiments bombarding gold 
foil with alpha particles. The gold foil experiment by Rutherford, first done in 1911, is of particular interest 
to us in this unit because it was this experiment that first gave science an approximate measurement for 
the size of the atom. He was able, using technology available in the early part of the 1900s, to determine 
quantitatively that the nucleus had an approximate size of 4 x 10 -12 cm. The size of the atom is slightly 
larger, approximately 2 X 10~ 8 cm in diameter. 

Atomic Size in a Column Increases from Top to Bottom 

Let's now look at the atomic radii or the size of the atom from the top of a family or group to the bottom. 
Take, for example, the Group 1 metals. Each atom in this family (and all other main group families) has 
the same number of electrons in the outer energy level as all the other atoms of that family. Each row 
(period) in the periodic table represents another added energy level. When we first learned about principal 
energy levels, we learned that each new energy level was larger than the one before. Energy level 2 is larger 
than energy level 1, energy level 3 is larger than energy level 2, and so on. Therefore, as we move down 
the periodic table from period to period, each successive period represents the addition of a larger energy 
level. It becomes apparent that as we move downward through a family of elements, that each new atom 
has added another energy level and will, therefore, be larger. 

Table 10.1: 

Element # of protons Electron Configuration # of energy levels 

Li 3 [He] 2s 1 2 

Na 11 [Ne] 3s 1 3 

K 19 [Ar] 4s l 4 

Rb 37 [Kr] 5s 1 5 

Cs 55 [Xe] 6s 1 6 

One other contributing factor to atomic size is something called the shielding effect. The protons in the 
nucleus attract the valence electrons in the outer energy level because of opposite electrostatic charges. 
The strength of this attraction depends on the size of the charges, the distance between the charges, AND 
the number of electrons in-between the nucleus and the valence electrons. The core electrons shield the 
valence electrons from the nucleus. The presence of the core electrons weakens the attraction between 
the nucleus and the valence electrons. This weakening of the attraction is called the shielding effect. The 
amount of shielding depends on the number of electrons between the nucleus and the valence electrons. 
The strength with which the nucleus pulls on the valence electrons can pull the valence shell in tighter 
when the attraction is strong and not so tight when the attraction is weakened. The more shielding that 
occurs, the further the valence shell can spread out. 

For example, if you are looking at the element sodium, it has the electron configuration: 
www.ckl2.org 312 



Na ls 2 2s 2 2p 6 3s 1 

The outer energy level is n = 3 and there is one valence electron but the attraction between this lone 
valence electron and the nucleus that has those 11 protons is shielded by the other 10 inner (or core) 
electrons. 

When we compare an atom of sodium to one of cesium, we notice that the number of protons increases 
as well as the number of energy levels occupied by electrons. There are also many more electrons between 
the outer electron and the nucleus, thereby shielding the attraction of the nucleus. 

Cs ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d 10 4p 6 5s 2 4d 10 5p 6 6s 1 

The outermost electron, 6s 1 , therefore, is held very loosely. In other words, because of shielding, the 
nucleus has less control over this 6s 1 electron than it does over a 3s 1 electron. The result of all of this is 
that the atom's size will be larger. Table 2 gives the values for the atomic radii for the group 1 metals plus 
a visual representation to appreciate the size change in a group in the periodic table. 

















4 


• Cs 






























< 


*Rb 












4 


.K 






















































































< 


'Na 


















































4 


, Li 














Atomic Radii 
(in pm) 

100- 








































































































50- 


























































































■ ■ 


1 1 1 1 1 1 



3 11 19 37 

Number of Protons in Nucleus 



Table 2: Atomic Radii Values for Group 1 Metal (measurement units for atomic radii are picometers 
(pm) or 1 X 10~ 12 meters) 

Table 10.2: 



Element 



Atomic radii 



Visual 



Li 



123 pm 



313 



www.ckl2.org 



Table 10.2: (continued) 



Element Atomic radii Visual 



Na 157 pm 



K 203 pm 



Rb 216 pm 



Cs 235 pm 





What is true for the Group I metals is true for all of the groups, or families, across the periodic table. 
As you move downward in the periodic table through a family group, the size of the atoms increases. For 
instance, the atoms that are the largest in the halogen family are bromine and iodine (since astatine is 
radioactive and only exists for short periods of time, we won't include it in the discussion). 

Sample question: Which of the following is larger? Explain. 

(a) As or Sb 

(b) Ca or Be 

(c) Polonium or Sulfur 
Solution: 

(a) Sb because it is below As in Group 15. 

(b) Ca because it is below Be in Group 2. 

(c) Po because it is below S in Group 16. 

As noted earlier for the main group metals, the outermost energy level in the electron configuration is 
indicated by the period number. So the element magnesium (Z = 12), is in period 3, group 2. According 
to this, we can say that there are 3 energy levels with 2 electrons in the outermost energy level. Let's look 
at the electron configuration for magnesium. 

Mg : ls 2 2s 2 2p 6 3s 2 

Moving from magnesium to strontium, strontium is in the 5th period of group 2. This means that there 
are two electrons in the valence energy level. Strontium also has electrons occupying five energy levels. 

Sr: ls 2 2s 2 2p 6 3s 2 3p 6 4s 2 3d l0 4p 6 5s 2 

You can imagine that with the increase in the number of energy levels, the size of the atom must increase. 
The increase in the number of energy levels in the electron cloud takes up more space. 



www.ckl2.org 314 



Therefore, the trend within a group or family on the periodic table is that the atomic size increases with 
increased number of energy levels. The periodic table below shows the trend of atomic size for groups. 
The arrow indicates the direction of the increase. 



(ft 

ca 
<u 

i— 
o 

c 





Atomic Size 




















■ 






























































































































. 






^H 











































































Atomic Size in a Period Decreases from Left to Right 

In order to determine the trend for the periods, we need to look at the number of protons (nuclear charge), 
the number of energy levels, and the shielding effect. For a row in the periodic table, the atomic number 
still increases (as it did for the groups) and thus the number of protons would increase. When we examine 
the energy levels for period 2, we find that the outermost energy level does not change as we increase the 
number of electrons. In period 2, each additional electron goes into the second energy level. So the number 
of energy levels does not go up. As we move from left to right across a period, the number of electrons in 
the outer energy level increases but it is the same outer energy level. Table 10.3 shows the electron 
configuration for the elements in period 2. 

Table 10.3: Electronic Configuration for Elements in Row 2 



Element 



# of protons 



Electron Configuration 



Lithium (Li) 
Beryllium (Be) 
Boron (B) 
Carbon (C) 
Nitrogen (N) 
Oxygen (O) 
Fluorine (F) 



3 

4 

5 
6 

7 



Is^s 1 

ls 2 2s 2 

ls 2 2s 2 2p 1 

ls 2 2s 2 2p 2 

ls 2 2s 2 2p 3 

ls 2 2s 2 2p* 

ls 2 2s 2 2p b 



Looking at the elements in period 2, the number of protons increases from lithium with three protons, to 
fluorine with nine protons. Therefore, the nuclear charge increases across a period. Meanwhile, the number 
of energy levels occupied by electrons remains the same. The numbers of electrons in the outermost 
energy level increases from left to right along a period but how will this affect the radius? We know 
that every one of the elements in row #2 has two electrons in their inner energy level (2 core electrons). 
The core electrons shield the outer electrons from the charge of the nucleus. With lithium, there are 
two core electrons and one valence electron so those two core electrons will shield the one outer electron. 
In beryllium, there are four protons being shielded by the Is 2 electrons. With the increasing number 
of protons attracting the outer electrons and the same shielding from the core electrons, the valence 



315 



www.ckl2.org 



electrons are pulled closer to the nucleus, making the atom smaller. 

In group 7A, the first element in the group is fluorine. With fluorine, there are 9 protons and 9 electrons. 
The electronic configuration is ls 2 2s 2 2p 5 . However, there are still the same core electrons as with lithium 
and beryllium, that is, the Is 2 electrons. Since there are more protons, there is an increase in the nuclear 
charge. With an increase in nuclear charge, there is an increase in the pull between the protons and the 
outer level, pulling the outer electrons toward the nucleus. The amount of shielding from the nucleus does 
not increase because the number of core electrons remains the same (Is 2 for this period). The net result is 
that the atomic size decreases going across the row. In Table 10.4, the values are shown for the atomic 
radii for the row starting at lithium and ending with fluorine plus a visual representation to appreciate the 
size change in a group in the periodic table. 

































































< 


,Li 




















< 


,Be 






















































Atomic Radii 






« 


>B 












(in pm) 








« 


'c . 


► N , 


>° , 


,F 




















4 


'tie 


















































































































1 1 1 



3453789 10 

Number of Protons in Nucleus 



Table 10.4: Atomic Radii Values for Period 2 



Element 



Atomic radii 



Visual 



Li 



123 pm 



Be 



111 pm 



B 



86 pm 



77 pm 




www.ckl2.org 



316 



Table 10.4: (continued) 



Element 



Atomic radii 



Visual 



N 



O 



74 pm 

73 pm 
72 pm 




o 

o 



Let's add this new trend to the periodic table. Look at the diagram below of our new "periodic trend 
table". In the diagram you will notice that the trend arrow for the period shows the atomic radii increase 
going from right to left, which is the same as decreasing from left to right. 



o 




Considering all the information about atomic size, you will recognize that the largest atom on the periodic 
table is all the way to the left and all the way to the bottom, francium, #87, and the smallest atom is all 
the way to the right and all the way to the top, helium, #2. 

The fact that the atoms get larger as you move downward in a family is probably exactly what you expected 
before you even read this section, but the fact that the atoms get smaller as you move to the right across 
a period is most likely a big surprise. Make sure you understand this trend and the reasons for it. 



For the Transition Elements, the Trend is Less Systematic 

A general trend for atomic radii in the periodic would look similar to that found in the diagram below. 
The elements with the smallest atomic radii are to the upper right; those with the largest atomic radii are 
to the lower left. 



317 



www.ckl2.org 



(A 
<D 
(A 
(d 
CD 
•— 
o 
c 















Increases 














Atomic Size 






























I 










^^ 


















- 




w^ 
























^T 











































































































Until now, we have worked solely with the main group metals. Let's look at our three factors and see how 
these factors fit the transition metal series. Looking at the first row of the transition metals, the 3d row, 
Table 10.5 shows the number of protons in each of the 10 elements in this row. The number of protons is 
increasing so the nuclear charge is increasing. 

Table 10.5: Summary of Data for 3d Metals 



Element 



# of protons 



Electron Configuration 



Scandium (Sc) 
Titanium (77) 
Vanadium (V) 
Chromium (Cr) 
Manganese (Mn) 
Iron (Fe) 
Cobalt (Co) 
Nickel (Ni) 
Copper (Cm) 
Zinc (Zn) 



21 
22 
23 
24 
25 
26 
27 
28 
29 
30 



[Ar 


3d l 4s 2 


[Ar 


3d 2 4s 2 


[Ar 


3d 3 4s 2 


[Ar 


3d b 4s 1 


[Ar 


3d 5 4s 2 


[Ar 


3d 6 4s 2 


[Ar 


3d 7 4s 2 


[Ar 


3d 8 4s 2 


[Ar 


3d w 4s 1 


[Ar 


3d w 4s 2 



The number of electrons are increasing, but in a particular way. We know that as the number of electrons 
increases going across a period, there is more pull of these electrons toward the nucleus. However, with 
the ^-electrons, there is some added electron-electron repulsion. Take a look at Table 10.5 and note the 
unusual electron configuration of chromium. 

In chromium, there is a promotion of one of the 4s electrons to half fill the 3d sublevel, the electron-electron 
repulsions are less and the atomic size is smaller. The opposite holds true for the latter part of the row. 
Table 10.6 shows the first row of the transition metals along with their size. 

Table 10.6: Atomic Radii for 3d Metals 



Element 



Atomic radii (pm) 



Scandium (Sc) 
Titanium (77) 
Vanadium (V) 
Chromium (Cr) 
Manganese (Mn) 



164 
147 
135 
129 
137 



www.ckl2.org 



318 



Table 10.6: (continued) 



Element 



Atomic radii (pm) 



Iron (Fe) 

Cobalt (Co) 
Nickel (Ni) 
Copper (Cu) 
Zinc (Zn) 



126 
125 
125 
128 
137 



The graph of the number of protons versus the atomic radii is shown below. 



200 



150 



atomic radii (pm) 



100 



50 



H 1 1 1 h 



-A 



H 1 1 1 1 1 1 1 1 1 1 1 h 



20 



25 3i 

Number of protons 



35 



Graphing the atomic number (or the number of protons) versus the atomic radii, we can see the trend in 
the 3d transition metals isn't quite as systematic as with the main group elements. 

Lesson Summary 

• Atomic size is the distance from the nucleus to the valence shell where the valence electrons are 
located. 

• The electrons surrounding the nucleus exist in an electron cloud. 

• You can predict the probability of where the electron is but not its exact location. 

• Atomic size is difficult to measure because it has no definite boundary. 

• Atomic radius is a more definite and measureable way of defining atomic size. It is the distance from 
the center of one atom to the center of another atom in a homonuclear diatomic molecule. 

• Rutherford led the way to determining the size of the atom with his gold foil experiment. 

• Mass spectrometers and other machines are available to measure masses and determine structure 
directly. 

• Atomic size is determined indirectly. 

• There are three factors that help in the prediction of the trends in the periodic table: number of 
protons in the nucleus, number of shells, and shielding effect. 



319 



www.ckl2.org 



• The atomic size increases from the top to the bottom in any group as a result of increases in all of 
the three factors. *As the number of energy levels increases, the size must increase. 

• Going across a period (from left to right), the number of protons increases and therefore the nuclear 
charge increases. *Going across a period, the number of electron energy levels remains the same but 
the number of electrons increases within these energy levels. Therefore the electrons are pulled in 
closer to the nucleus. 

• Shielding is relatively constant since the core electrons remain the same. 

• The trend in the periodic table is that as you move across the periodic table from left to right, the 
atomic radii decrease. This trend is not as systematic for the transition metals because other factors 
come into play. 

Review Questions 

1. Why is the atomic size considered to have "no definite boundary"? 

2. How is atomic size measured? 

(a) using a spectrophotomer 

(b) using a tiny ruler (called a nano ruler) 

(c) indirectly 

(d) directly 

3. Draw a visual representation of the atomic radii of an iodine molecule. 

4. Which of the following would be smaller? 

(a) In or Ga 

(b) K or Cs 

(c) Te or Po 

5. Explain in your own words why Iodine is larger than Bromine. 

6. What three factors determine the trend of atomic size going down a group? 

7. What groups tend to show this trend? 

8. Which of the following would have the largest atomic radii? 

(a) Si 

(b) C 

(c) Sn 

(d) Pb 

9. Which of the following would have the smallest atomic radius? 

(a) 2s 2 

(b) 4s 2 4p 3 

(c) 2s 2 2p A 

(d) As 2 

10. Arrange the following in order of increasing atomic radii: Tl,B,Ga,Al,In. 

11. Arrange the following in order of increasing atomic radii: Ge,Sn,C, 

12. Which of the following would be larger? 

(a) Rb or Sn 

(b) Ca or As 

13. Place the following in order of increasing atomic radii: Mg,Cl,S,Na. 

14. Describe the atomic size trend for the rows in the periodic table. 

15. Draw a visual representation of the periodic table describing the trend of atomic size. 

16. Which of the following would have the largest atomic radii? 

www.ckl2.org 320 



(a) 


Sr 


(b) 


Sn 


(c) 


Rb 


(d) 

M "hi 


In 


>\ ill 

(a) 


en C 

K 


(b) 


Kr 


(c) 


Ga 


(d) 


Ge 



17. Which of the following would have the smallest atomic radii? 



18. Place the following elements in order of increasing atomic radii: In, Ca, Mg, Sb, Xe. 

19. Place the following elements in order of decreasing atomic radii: Al,Ge,Sr,Bi,Cs. 

20. Knowing the trend for the rows, what would you predict to be the effect on the atomic radius if an 
atom were to gain an electron? Use an example in your explanation. 

21. Knowing the trend for the rows, what would you predict to be the effect on the atomic radius if the 
atom were to lose an electron? Use an example in your explanation. 

Vocabulary 

atomic size Atomic size is the distance from the nucleus to the valence shell where the valence electrons 
are located. 

atomic radius One-half the distance between the centers of the two atoms of a homonuclear molecule. 

nuclear charge The number of protons in the nucleus. 

shielding effect The core electrons in an atom interfere with the attraction of the nucleus for the out- 
ermost electrons. 

electron-electron repulsion The separation that occurs because electrons have the same charge. 



10.2 Ionization Energy 

Lesson Objectives 



Define ionization energy. 

Describe the trend that exists in the periodic table for ionization energy. 

Describe the ionic size trend that exists when elements lose one electron. 



Introduction 

When we study the trends in the periodic table, we cannot stop at just atomic size. In this section of the 
chapter, we will begin an understanding of an important concept, namely ionization energy and recognize 
its trend on the periodic table. 



321 www.ckl2.org 



Ionization Energy is the Energy Required to Remove an Electron 



Lithium has an electron configuration of ls 2 2s 1 . Lithium has one electron in its outermost energy level. 
In order to remove this electron, energy must be added to the system. Look at the equation below: 



energy + L/ (g) 
ls 2 2s l 



Li (g) + e 
Is 2 



With the addition of energy, a lithium ion can be formed from the lithium atom by losing one electron. 
This energy is known as the ionization energy. The ionization energy is the energy required to remove 
the most loosely held electron from a gaseous atom or ion. "In the gaseous phase" is specified because in 
liquid or solid, other energies get involved. The general equation for the ionization energy is as follows. 



energy + A fe) 



A Z) +e ~ 



The higher the value of the ionization energy, the harder it is to remove that electron. We can see a trend 
when we look at the ionization energies for the elements in period 2. Table 10.7 summarizes the electron 
configuration and the ionization energies for the elements in the second period. 

Table 10.7: First Ionization Energies 



Element 



Electron configuration 



First Ionization Energy, IE\ 



Lithium (Li) 
Beryllium (Be) 
Boron (B) 
Carbon (C) 
Nitrogen (N) 
Oxygen (O) 
Fluorine (F) 



[He] 


2s' 


[He] 


2s 2 


[He] 


2s 2 2p l 


[He] 


2s 2 2p 2 


[He] 


2s 2 2p 3 


[He] 


2s 2 2p A 


[He] 


2s 2 2p 5 



520 kJ/mol 
899 kJ/mol 
801 kJ/mol 
1086 kJ/mol 
1400 kJ/mol 
1314 kJ/mol 
1680 kJ/mol 



When we look closely at the data presented in Table 10.7, we can see that as we move across the period 
from left to right, in general, the ionization energy increases. At the beginning of the period, with the 
alkali metals and the alkaline earth metals, losing one or two electrons allows these atoms to become ions. 

energy + Li(g) -> Li + (g) + e~ 
energy + Mg(g) -» Mg 2+ + 2e~ 

[He] 2s 1 -> [He] + e~ 
[Ne] 3s 2 -> [Ne] + 2e~ 

As we move across the period, the atoms become smaller which causes the nucleus to have greater attraction 
for the valence electrons. Therefore, the electrons are more difficult to remove. 

A similar trend can be seen for the elements within a family. Table 10.8 shows the electron configuration 
and the first ionization energies (IE\) for some of the elements in the first group, the alkali metals. 



www.ckl2.org 



322 



Table 10.8: Ionization Energies for Some Group 1 Elements 



Element 



Electron configuration 



First Ionization Energy, IE\ 



Lithium (Li) 
Sodium (No) 
Potassium (K) 



[He] 2s 1 
[He] 3s 1 
[He] 4s 1 



520 kJ/mol 
495.5 kJ/mol 
418.7 kJ/mol 



Comparing the electron configurations of lithium to potassium, we know that the electron to be removed 
is further away from the nucleus. We know this because the n value is larger meaning the energy level 
where the valence electron is held is larger. Therefore it is easier to remove the most loosely held electron 
because the atom is larger with a greater shielding effect which means that the nucleus has less control 
over potassium's outer electron, 4s 1 . 

Therefore IE\ for potassium (418.7 kJ/mol) is less than IE\ for lithium (520 kJ/mol). 

If a second electron is to be removed from an atom, the general equations are the following: 



/(g) + energy -> / 1+ + e 
J+ (g) + ener gy ~> j2+ (g) + e ~ 



IE l 
IE 2 



Since there is an imbalance of positive and negative charges when a second electron is being removed, 
the energy required for the second ionization (IE2) will be greater than the energy required for the first 
ionization (IE\). Simply put, IE\ < IE2 < IE3 < IE4. 

The Charge on the Nucleus Increases and Size Decreases 

So if we look at the ionization energy trend in the periodic table and add it to the trend that exists with 
atomic size we can show the following on our periodic table chart. 







Ionization Energy Increases 








Atomic Size Increases 































































































































































































































But why does the ionization energy increase going across a period and decrease going down a group? It has 
to do with two factors. One factor is that the atomic size decreases. The second factor is that the effective 
nuclear charge increases. The effective nuclear charge is the charge experienced by a specific electron 
within an atom. Remember, the nuclear charge was used to describe why the atomic size decreased going 
across a period. When we look at the data in Table 10.7 again, we can see how the effective nuclear charge 
increases going across a period. Table 10.9 shows the effective nuclear charge along with the ionization 
energy for the elements in period 2. 



323 



www.cki2.0rg 



Table 10.9: Effective Nuclear Charge for Period 2 Main Group Elements 



Element 



Electron con- 
figuration 



# of protons # of core elec- Effective nu- Ionization En- 
trons clear charge ergy 



Lithium (Li) 


[He] 


2s 1 


3 


Beryllium (Be) 


[He] 


2s 2 


4 


Boron (B) 


[He] 


2s 2 2p 1 


5 


Carbon (C) 


[He] 


2s 2 2p 2 


6 


Nitrogen (N) 


[He] 


2s 2 2p 3 


7 


Oxygen (0) 


[He] 


2s 2 2p 4 


8 


Fluorine (F) 


[He] 


2s 2 2p 5 


9 



2 
2 
2 
2 
2 
2 
2 



1 
2 
3 

4 
5 
6 
7 



520 kJ/mol 
899 kJ/mol 
801 kJ/mol 
1086 kJ/mol 
1400 kJ/mol 
1314 kJ/mol 
1680 kJ/mol 



The electrons that are shielding the nuclear charge are the core electrons, or in the case of period 2, the Is 2 
electrons. The effective nuclear charge is the difference between the total charge in the nucleus (the number 
of protons) and the number of shielded electrons. Notice how as the effective nuclear charge increases, so 
does the ionization energy. Overall the general trend for ionization energy is shown below. 

Ionization Energy 



H 


1 








Inert 


;ase< 






~ 


IB 

a - 
<u 

u 

a 

i— i - 
























l^n 





























































































Sample question: 

What would be the effective nuclear charge for CI? Would you predict the ionization energy to be higher 
or lower than fluorine? 

Solution: 

Chlorine has the electronic configuration of: 

CI : [Ne] 3s 2 3p 5 

The effective nuclear charge is 7, the same as fluorine. Predicting the ionization energy would be difficult 
with this alone. The atomic size, however, is larger for chlorine than for fluorine because now there are 
three energy levels (chlorine is in period 3). Now we can say that the ionization energy should be lower 
than that of fluorine because the electron would be easier to pull off the level further away from the nucleus. 
(Indeed, the value for chlorine is 1251 kJ/mol). 

Some Anomalies With the Trend in Ionization Energy 

There are a few anomalies that exist with respect to the ionization energy trends. Going across a period 
there are two ways in which the ionization energy may be affected by the electron configuration. When we 
look at period 3, we can see that there is an anomaly observed as we move from the 3s sublevels to the 3p 



www.ckl2.org 



324 



sublevel. The table below shows the electron configurations for the main group elements in period 3 along 
with the first ionization energy for these elements. 



Table 4: Ionization Energies for Period 3 Main Group Elements 


Element 


Electron Configuration 


Ionization Energy 


Sodium (Na) 


1s 2 2s 2 2p e 3s 1 


495.9 kJ/mol 


Magnesium (Mg} 


1s 2 2s 2 2p 6 3s 2 


738.1 kJ/mol 1 


Aluminum (Al) 


1s 2 2s 2 2p 6 3s 2 3p 1 


577.9 kJ/mol J 


Silicon [Si) 


1s 2 2s 2 2p 6 3s 2 3p 2 


786.3 kJ/mol 


Phosphorus [P) 


1s 2 2s 2 2p 6 3s 2 3p 3 


1012 kJ/mol 


Sulfur [S) 


1s 2 2s 2 2p 6 3s 2 3p 4 


999.5 kJ/mol 


Chlorine [CI) 


1s 2 2s 2 2p 6 3s 2 3p 5 


1251 kJ/mol 


Argon (Ar) 


1s 2 2s 2 2p 6 3s 2 3p 6 


1520 kJ/mol 



3s 



f| 3p 



In the table we see that when we compare magnesium to aluminum the first IE decreases instead of 
increasing as we would have expected. So why would this be so? Magnesium has its outermost electrons 
in the s sublevel. The aluminum atom has its outermost electron in the 2>p sublevel. Since p electrons 
have just slightly more energy than s electrons, it takes a little less energy to remove that electron from 
aluminum. One other slight factor is that the electrons in 3s 2 shield the electron in 3p . These two factors 
allow the IE\ for Al to be less than IE\ for Mg. 

When we look again at Table 1 below, we can see that the ionization energy for nitrogen seems out of 
place. 



Table 1 : Ionization Energies for Period 2 Main Group Elements 




Element 


Electron Configuration 


Ionization Energy 




Lithium [Li) 


[He]2s 1 


520 kJ/mol 




Beryllium [Be) 


[He] 2s 2 


899 kJ/mol 




Boron [B) 


[He] 2s 2 2p 1 


801 kJ/mol 




Carbon [C) 


[He] 2s 2 2p 2 


1086 kJ/mol 


1 


Nitrogen (N) 


[He]2s 2 2p 3 


1400 kJ/mol 4 


{ 


Oxygen (0) 


[He] 2s 2 2p 4 




I 


_ 
1314kJ/mol 




Fluorine (F) 


[He]2s 2 2p" 


1680kJ/mol 




Neon (Ne) 


[He] 2s 2 2p 6 


2081 kJ/mol 





_t_ _L_t_ 

\\ 2P 

2s 

HJ_i 

2s 



While nitrogen has one electron occupying each of the three p orbitals in the 2 nd sublevel, oxygen has 



325 



www.ckl2.org 



two orbitals occupied by only one electron but one orbital containing a pair of electrons. The greater 
electron-electron repulsion experienced by these 2p electrons allows for less energy to be needed to remove 
one of these. Therefore, IE\ for oxygen is less than nitrogen. 

Lesson Summary 

• Ionization energy is the energy required to remove the most loosely held electron from a gaseous 
atom or ion. Ionization energy generally increases across a period and decreases down a group. The 
effective nuclear charge is the charge of the nucleus felt by the valence electron. 

• The effective nuclear charge and the atomic size help explain the trend of ionization energy. Going 
down a group the atomic size gets larger and the electrons can be more readily removed, therefore, 
ionization energy decreases. Going across a period the effective nuclear charge increase so the electrons 
are harder to remove and the ionization energy increases. Once one electron has been removed, a 
second electron can be removed but IE\ < IE<i- If a third electron is removed, IE\ < IE2 < IE3 and 
so on. 

Review Questions 

1. Define ionization energy and show an example ionization equation. 

2. Draw a visual representation of the periodic table describing the trend of ionization energy. 

3. Which of the following would have the largest ionization energy? 

(a) Na 

(b) Al 

(c) H 

(d) He 

4. Which of the following would have the smallest ionization energy? 

(a) K 

(b) P 

(c) S 

(d) Ca 

5. Place the following elements in order of increasing ionization energy: Na,0,Ca,Ne,K. 

6. Place the following elements in order of decreasing ionization energy: N,Si,S,Mg,He. 

7. Using experimental data, the first ionization energy for an element was found to be 600 kJ/mol. The 
second ionization energy for the ion formed was found to be 1800 kJ/mol. The third ionization energy 
for the ion formed was found to be 2700 kJ/mol. The fourth ionization energy for the ion formed 
was found to be 11600 kJ/mol. And finally the fifth ionization energy was found to be 15000 kJ/mol. 
Write the reactions for the data represented in this question. Which group does this element belong? 
Explain. 

8. Using electron configurations and your understanding of ionization energy, which would you predict 
would have higher second ionization energy: Na or Mg? 

9. Comparing the first ionization energy {IE\) of calcium, Ca, and magnesium, Mg, : 

(a) Ca has a higher IE\ because its radius is smaller. 

(b) Mg has a higher IE\ because its radius is smaller. 

(c) Ca has a higher IE\ because its outer sub-shell is full. 

(d) Mg has a higher IE\ because its outer sub-shell is full. 

(e) they have the same IE\ because they have the same number of valence electrons. 

10. Comparing the first ionization energy {IE\) of beryllium, Be, and boron, B, : 
www.ckl2.org 326 



(a) Be has a higher IE\ because its radius is smaller. 

(b) B has a higher IE\ because its radius is smaller. 

(c) Be has a higher IE\ because its s sub-shell is full. 

(d) B has a higher IE\ because its s sub-shell is full. 

(e) They have the same IE\ because B has only one more electron than Be. 

Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

ionization energy The energy required to remove an electron from a gaseous atom or ion: 

energy + J(g) — > J + (g) + e~ (first ionization energy). 

effective nuclear charge The charge on the atom or ion felt by the outermost electrons (valence elec- 
trons). 

10.3 Electron Affinity 

Lesson Objectives 

• Define electron affinity. 

• Describe the trend for electron affinity on the periodic table. 

Introduction 

The final periodic trend for our discussion is electron affinity. We have talked about atomic structure, 
electronic configurations, size of the atoms and ionization energy. And now, the final periodic trend we 
will study is how an atom can gain an electron and the trends that exist in the periodic table. 

The Energy Process When an Electron is Added to an Atom 

Atoms can gain or lose electrons. When an atom gains an electron, energy is given off and is known as 
the electron affinity. Electron affinity is defined as the energy released when an electron is added to a 
gaseous atom or ion. 

T(g)+e- ^ T- [g) 

When most reactions occur that involve the addition of an electron to a gaseous atom, potential energy is 
released. 

Br {g) + e~ -> Br l ~( g) EA = -325 kJ/mol 

[Ar] 4s 2 4p 5 [Ar] 4s 2 4p 6 

Let's look at the electron configurations of a few of the elements and the trend that develops within groups 
and periods. Take a look at Table 10.10, the electron affinity for the Halogen family. 

327 www.ckl2.org 



Table 10.10: Electron Affinities for Group 7A 



Element 



Electron configuration 



Electron affinity, kJ/mol 



Fluorine (F) 
Chlorine (CI) 
Bromine (Br) 
Iodine (7) 



[He] 2s 2 2p 5 
[Ne] 3s 2 3p 5 
[Ar] 4s 2 4p 5 
[Kr] 5s 2 5p 5 



-328 
-349 
-325 
-295 



As you can see, the electron affinity generally decreases (becomes less negative) going down a group because 
of the increase in size of the atoms. Remember that the atoms located within a family but lower on the 
periodic table are larger since there are more electrons filling more energy levels. For example, an atom of 
chlorine is smaller than iodine; or, an atom of oxygen is smaller than sulfur. When an electron is added to 
a large atom, less energy is released because the electron cannot move as close to the nucleus as it can in 
a smaller atom. Therefore, as the atoms in a family get larger, the electron affinity gets smaller. 

There is an exception to this when it involves certain small atoms. Electron affinity for fluorine is less than 
chlorine most likely due to the electron-electron repulsions that occur between the electrons where n = 2. 
This phenomenon is observed in other families as well. For instance, the electron affinity for oxygen is 
less than the electron affinity for sulfur. Electron affinity of all of the elements in the second period is less 
than the ones below them due to the fact that the elements in the second period have such small electron 
clouds that electron repulsion is greater than that of the rest of the family. 

Nonmetals Tend to Have the Highest Electron Affinity 

Overall, the periodic table shows the general trend similar to the one below. 



Table 5.1: Electron Affinities for Period 4 Main Group Elements 




Element 


Electron Configuration 


Electron Affinity 




Potassium (K) 


[Ar] 4s 1 


-48 kJ/mol 




Calcium (Ca) 


[Ar]4s 2 






-2.4 kJ/mol H__^ 






Gallium (Ga) 


[Ar] 4s 2 4p 1 


-29 kJ/mol 




Germanium (Ge) 


[Ar]4s 2 4p 2 


-118 kJ/mol 




Arsenic (As) 


[Ar]4s 2 4p 3 






-77 kJ/mol E_ 






Selenium (Se) 


[Ar]4s 2 4p 4 


-195 kJ/mol 




Bromine (Br) 


[Ar]4s 2 4p 5 


-325 kJ/mol 




Krypton [Kr) 


[Ar] 4s 2 4p 6 


kJ/mol 






4p 



4s 



LJ_i 

H 



4s 



The general trend in the electron affinity for atoms is almost the same as the trend for ionization energy. 
This is because both electron affinity and ionization energy are highly related to atomic size. Large atoms 
have low ionization energy and low electron affinity. Therefore, they tend to lose electrons and do not 
tend to gain electrons. Small atoms, in general, are the opposite. Since they are small, they have high 
ionization energies and high electron affinities. Therefore, the small atoms tend to gain electrons and tend 
not to lose electrons. The major exception to this rule are the noble gases. They are small atoms and 



www.ckl2.org 



328 



do follow the general trend for ionization energies. The noble gases, however, do not follow the general 
trend for electron affinities. Even though the noble gases are small atoms, their outer energy levels are 
completely filled with electrons and therefore, an added electron cannot enter their outer most energy level. 
Any electrons added to a noble gas would have to be the first electron in a new (larger) energy level. This 
causes the noble gases to have essentially zero electron affinity. This concept is discussed more thoroughly 
in the next chapter. 

When atoms become ions, the process involves either the energy released through electron affinity or 
energy being absorbed with ionization energy. Therefore, the atoms that require a large amount of energy 
to release an electron will most likely be the atoms that give off the most energy while accepting an 
electron. In other words, non-metals will most easily gain electrons since they have large electron affinities 
and large ionization energies; and, metals will lose electrons since they have the low ionization energies 
and low electron affinities. 

Now let's add this last periodic trend to our periodic table representation and our periodic trends are 
complete. 



Atomic Size Increases 



Ionization Energy Increases 





* * 






Electron Affinity 1 


icreases 






IS, 

m 

IS, 


0) 
is, 

to 
b 

i 

LU 
o 

£Z 

r s 


1 ai 

cu 
is, 
ra 
a> 

b 

ez 
% 

5 

T3 
a) 

LU 


\ 





















r 












b 


















s 




— 


































to 


































E 




































< 






























1 
































T 































Note: Both the trend in atomic size and 
the trend in ionization energy extend all 
the way left to right and top to bottom in 
the periodic table, but the trend in electron 
affinity only extends through family 7A, the 
halogens. The noble gases, family 8A, 
essentially have zero electron affinity. 



The web site below contains some textual material discussing chemical bonds and also allows you to access 
tables of graphs showing the ionization energies and electron affinities of various elements. Chemical Bond 
Data (http : //hyperphysics . phy-astr . gsu . edu/hbase/chemical/bondd . html) 

The development and arrangement of the periodic table of elements is examined. Video on Demand - The 
World of Chemistry - The Periodic Table (http: //www. learner. org/vod/vod_window.html?pid=799) 

Lesson Summary 

• Electron affinity is the energy required (or released) when an electron is added to a gaseous atom or 
ion. Electron affinity generally increases going up a group and increases left to right across a period. 

• Non-metals tend to have the highest electron affinities. 

Review Questions 

1. Define electron affinity and show an example equation. 

2. Choose the element in each pair that has the lower electron affinity: 

(a) Li or N 



329 



www.ckl2.org 



(b) Na or CI 

(c) Ca or K 

(d) Mg or F 

3. Why is the electron affinity for calcium much higher than that of potassium? 

4. Draw a visual representation of the periodic table describing the trend of electron affinity. 

5. Which of the following would have the largest electron affinity? 

(a) Se 

(b) F 

(c) Ne 

(d) Br 

6. Which of the following would have the smallest electron affinity? 

(a) Na 

(b) Ne 

(c) Al 

(d) /?£> 

7. Place the following elements in order of increasing electron affinity: Te,Br,S,K,Ar. 

8. Place the following elements in order of decreasing electron affinity: S,Sn,Pb,F,Cs. 

9. Describe the trend that would occur for electron affinities for elements in period 3. Are there any 
anomalies? Explain. 

10. Comparing the electron affinity (EA) of sulfur, S , and phosphorus, P, : 

(a) S has a higher EA because its radius is smaller. 

(b) P has a higher EA because its radius is smaller. 

(c) S has a higher EA because its p sub-shell is half full. 

(d) P has a higher EA because its p sub-shell is half full. 

(e) they have the same EA because they are next to each other in the periodic table. 

Vocabulary 

electron affinity The energy required to add an electron to a gaseous atom or ion. 

T(g) + e- -» T-(g) 



Image Sources 



www.ckl2.org 330 



Chapter 11 
Covalent Bonding 



11.1 The Covalent Bond 

Lesson Objectives 

• The student will describe the basic nature of covalent bond formation. 

• The student will explain the difference between ionic and covalent bonding. 

• The student will state the relationship between molecular stability and bond strength. 

Introduction 

In ionic bonding, electrons leave metallic atoms and enter non-metallic atoms. This complete transfer 
of electrons changes both of the atoms into ions. Often, however, two atoms combine in a way that no 
complete transfer of electrons occurs. Instead, electrons are held in overlapping orbitals of the two atoms, 
so that the atoms are sharing the electrons. The shared electrons occupy the valence orbitals of both atoms 
at the same time. The nuclei of both atoms are attracted to this shared pair of electrons and the atoms 
are held together by this attractive force. The attractive force produced by sharing electrons is called a 
covalent bond. 

Sharing Electrons 

In covalent bonding, the atoms acquire a stable octet of electrons by sharing electrons. The covalent 
bonding process produces molecular substances as opposed to the lattice structures of ionic bonding. The 
covalent bond, in general, is much stronger than ionic bonds and there are far more covalently bonded 
substances than ionic substances. 

The diatomic hydrogen molecule, H2, is one of the many molecules that are covalently bonded. Each 
hydrogen atom has a Is electron cloud containing one electron. These Is electron clouds overlap and 
produce a common volume which the two electrons occupy (Figure 11.1). 

The shared pair of electrons spend more time between the two atoms than they do in other parts of the 
two Is orbitals but the shared pair of electrons occupies all of the Is orbitals of both atoms. 

In the simulated probability pattern for the overlapped Is orbitals in an H2 molecule, we see the electrons 
are still more likely to be found close to the nucleus than far away, but we also see that they spend more 
time between the two nuclei than they do on the far sides of the atoms. The extra time spent between the 

331 www.ckl2.org 



Figure 11.1: The orbitals of two hydrogen atoms overlap and share the two valence electrons. 



two nuclei is the source of the attraction that holds the atoms together in a covalent bond (Figure 11.2). 




Figure 11.2: Simulated probability pattern for the overlapped orbitals in . 

The diatomic fluorine molecule is also a covalently bonded atom. In the case of fluorine atoms, the atoms 
have filled Is orbitals, filled 2s orbitals, and two of the three 2p orbitals are full. Each atom has a half-filled 
2p orbital that is available to be overlapped. 




Figure 11.3: Showing the orbitals of fluorine with two orbitals full and one orbital half- full and then showing 
the half-filled orbital of each atom overlapping. 



Figure 11.3 shows the available (half-filled) 2p orbitals of two fluorine atoms overlapping to form a covalent 
bond. We use several methods to represent a covalent bond so that we don't have to spend all our time 
drawing orbitals. We can represent the bond in the F2 molecule with an electron dot formula (Figure 
11.4). 



www.ckl2.org 



332 



:f:f: :f-F 



o 



• • 



Figure 11.4: We can also show a covalent bond between atoms with an electron dot formula where the 
shared pair of electrons are the bonding electrons or with the bond represented by a dash. 



Bond Strength 

When atoms that attract each other move closer together, the potential energy of the system (the two 
atoms) decreases. When a covalent bond is formed, the atoms move so close together that the atoms 
overlap their electron clouds. As the atoms continue to move closer yet, the potential energy of the 
system continues to decrease ... to a point. If you continue to move atoms closer and closer together, 
eventually the two nuclei will begin to repel each other. If you push the nuclei closer together at this 
point, the repulsion causes the potential energy to increase. Each pair of covalently bonding atoms will 
have a distance between their nuclei that is the lowest potential energy distance. This position has the 
atoms close enough that the attraction between the nucleus of each one and the electrons of the other is 
maximum but the nuclei have not begun to repel each other strongly. For bonding atoms, this distance 
occurs somewhere after the electron clouds have overlapped. At this position, the atoms are at lowest 
potential energy. If the atoms are pushed closer, the potential energy goes up because you are crowding 
the nuclei together and if the atoms are pulled apart, potential energy goes up because you are separating 
particles that attract each other. Since this is the lowest potential energy position, the atoms will remain 
at the distance, bonded together. This distance is called the bond length. The more potential energy that 
was given up as this bond formed, the stronger the bond will be. If you want to break this bond, you must 
input all the potential energy that was given up as the bond was formed. 

The strength of a diatomic covalent bond can be expressed by the amount of energy necessary to break the 
bond and produce separate atoms. The energy needed to break a covalent bond is called bond energy 
and is measured in kilojoules per mole. Bond energy is used as a measure of bond strength. The bond 
strength of HBr is 365 kilojoules per mole. i.e. It would take 365 kilojoules to break all the chemical bonds 
in 6.02 x 10 23 molecules of HBr and produce separate hydrogen and bromine atoms. The bond strength of 
HCl is 431 kilojoules per mole. Consequently, the bond in HCl is stronger than the bond in HBr. 



Molecular Stability 

Compounds with high bond strengths are difficult to break up and therefore are stable compounds. When 
stable compounds are formed, large amounts of energy are given off so these molecules are in a relatively 
low energy state. In order to break the molecule apart, all the energy that was given off must be put back 
in. Low energy state bonds are stable and have high bond strength. 

Molecules with high bond energy have weak bonds. They did not release much energy when they formed 
and so not much energy is needed to the break the molecules back apart. High bond energy means low 
bond strength. 

Atoms of carbon, hydrogen, and oxygen can be chemically bonded in more than one way. In the molecule 
shown below, these atoms are bonded in a way that produces a molecule of glucose. 

333 www.ckl2.org 



CHO 

H-C-OH 

I 
HO-C-H 

I 
H-C-OH 

I 
H-C-OH 

I 

CHiOH 



+ 6 2 — *►" 6 COi + 6 H 2 + ENERGY 



The molecule of glucose can be reacted with six oxygen atoms to produce six molecules of carbon dioxide 
and six molecules of water. During the reaction, the atoms of the glucose molecule are rearranged into the 
structure of carbon dioxide and water molecules. The bonds in the glucose are broken and new bonds are 
formed. As this occurs, potential energy is released because the new bonds have lower potential energy 
that the original bonds. The bonds in the products are lower energy bonds and therefore, the product 
molecules are more stable. 

Some Compounds Have Both Covalent and Ionic Bonds 

If you recall the introduction of polyatomic ions, you will remember that the bonds that hold the polyatomic 
ions together are covalent bonds. Once the polyatomic ion is constructed with covalent bonds, it reacts 
with other substances as an ion. The bond between a polyatomic ion and another ion will be ionic. An 
example of this type of situation is in the compound sodium nitrate. Sodium nitrate is composed of a 
sodium ion and a nitrate ion. The nitrate ion is held together by covalent bonds and the nitrate ion is 
attached to the sodium ion by an ionic bond (Figure 11.5). 

covalent 

NaN0 3 




covalent 
ionic 



Figure 11.5: The bonding in sodium nitrate, . 

Lesson Summary 

• Covalent bonds are formed by electrons being shared between two atoms. 

• Half-filled orbitals of two atoms are overlapped and the valence electrons shared by the atoms. 
www.ckl2.org 334 



• Bond energy is the amount of energy necessary to break the covalent bond. 

• The strength of a covalent bond is measured by the bond energy. 

• Stable compounds have high bond energy and unstable compounds have low bond energy. 

Review Questions 

1. Describe the characteristics of two atoms that would be expected to form an ionic bond. 

2. Describe the characteristics of two atoms that would be expected to form a covalent bond. 

3. If an atom had a very high bond energy, would you expect it to be stable or unstable? 

4. When gaseous potassium ions and gaseous fluoride ions join together to form a crystal lattice, the 
amount of energy released is 821 kJ/mol. When gaseous potassium ions and gaseous chloride ions 
join together to form a crystal lattice, the amount of energy released is 715 kJ/mol. Which is the 
stronger bond, KF or KCll It these two compounds were increasingly heated, which compound would 
break apart at the lower temperature? 

Further Reading / Supplemental Links 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 4-3 is on Ionic vs. Covalent Bonds. 

• http : //www . f ordhamprep . org/gcurran/sho/sho/lessons/lesson43 . htm 

Vocabulary 

covalent bond A type of chemical bond where two atoms are connected to each other by the sharing of 
two or more electrons in overlapped orbitals. 

covalent bond strength The strength of a covalent bond is measured by the amount of energy required 
to break the bond. 

11.2 Atoms that Form Covalent Bonds 

Lesson Objectives 

• The student identify pairs of atoms that will form covalent bonds. 

• The student will draw Lewis structures for simple covalent molecules. 

• The student will identify sigma and pi bonds in a Lewis structure. 

Introduction 

A great deal of importance seems to have been attached to the electron configurations of noble gases or to 
the octet of electrons. These structures have been made in some way to seem to be a "desirable" thing for 
an atom to have. It is hoped that you do not see atoms and electrons as behaving the way they do because 

335 www.ckl2.org 



of "wants" and "desires". The bonding of atoms is directly by the laws of nature relating to the tendency 
toward minimum potential energy, electrical attraction and repulsion, and the arrangement of electrons 
in atoms. As it happens, these laws of nature and energy conditions do favor (in most cases) an octet of 
electrons for atoms. In ionic bonding, the atoms acquired this octet by gaining or losing electrons and, 
in covalent bonding, as you have seen, the atoms acquire the noble gas electron configuration by sharing 
electrons. 

Non-Metals Bond with Non- Metals to Form Covalent Bonds 

In the discussion of ionic bonds, it was clear that ionic bonds form between metals and non-metals because 
the high electron affinity non-metals are able to take electrons away from metals. Metals lose their electrons 
readily and have no attraction to add electrons. Since covalent bonds require that electrons be shared, 
it becomes apparent that metals will form few if any covalent bonds. Metals simply do not hold on to 
electrons with enough strength to form much in the way of covalent bonds. For a covalent bond to form, 
we need two atoms that both attract electrons with high electron affinity. Hence, the great majority of 
covalent bonds will be formed between two non-metals. When both atoms in a bond are from the right 
side of the periodic table, you can be sure that the bond is covalent. Here are some examples of molecules 
with covalent bonds. You should check to see where the atoms involved in these molecules appear in the 
periodic table. Covalent bonds form between atoms with relatively high electron affinity and they form 
individual, separate molecules (Figure 11.6). 



Phosphorus Trichloride 

PCI, 




CI 



CI 



:ci:P:ci: 
":Cl: " 

• • 

:CI-P-Cl: 
:CI: 



H 2 




(o 


) 


«v< 


H 


H 
:o:H 

• • 




^ 




:0-H 





Carbon Dioxide 

co„ 




:0:: C::o: 



:o = C = o: 



Figure 11.6: Various methods of showing a covalent bond. 

The differences between ionic and covalent bonds are explained by the use of scientific models and examples 
from nature. Video on Demand - The World of Chemistry - Chemical Bonds (http : //www . learner . org/ 
vod/vod_window . html?pid=800) 

Multiple Bonds 

So far, we have discussed covalent bonds in which one pair of electrons is shared. This type of bond is 
called a single bond. Some atoms can share more than one pair of electrons. When atoms share two pairs 
of electrons, it is called a double bond and when atoms share three pairs of electrons, it is called a triple 
bond. 

Double bonds are formed when atoms overlap two orbitals at the same time. An example of the formation 



www.ckl2.org 



336 



of a double bond is in the bonding in the O2 molecule. Oxygen has six valence electrons and they are 
distributed in the outermost energy level as 2s 2 2p 2 x 2p\2pl. Two of the p orbitals in the bonding shell of 
oxygen are only half-filled and therefore are available for overlap. One of these half-filled p orbitals from 
each oxygen atom will overlap end-to-end as shown in Figure 11.7. 



Figure 11.7: The end-to-end overlap of orbitals forms a sigma bond. 

Bonds formed by end-to-end overlap are called sigma (cr) bonds. The first bond formed between two atoms 
is always a sigma bond. The second half-filled orbitals of the oxygen atoms will be oriented vertically and 
cannot overlap end-to-end but they can overlap side-to-side as shown in the figure below. 

The side-to-side overlap of orbitals forms a pi bond. (Source: CK-12 Foundation. CC-BY-SA) 




The side-to-side overlap of orbitals forms a bond called a pi (n) bond. All the bonds formed between two 
atoms after the first bond are pi bonds. The electron dot formula for a double bond would show two pairs 
of electrons shared between atoms (see below figure) . 

Oxygen is double bonded. (Source: Richard Parsons. CC-BY-SA) 



Double bonds are stronger than single bonds. They are not exactly twice as strong; sometimes they are 
more than twice as strong and sometimes they are less than twice as strong. Oxygen is a reactive element 
and it is surprising that there is so much elemental oxygen in our atmosphere. The explanation for the 
existence of so much elemental oxygen is that the double bond is very strong and it takes a great deal of 
energy to break the double bond in oxygen so that the oxygen atoms could react with something else. 

The nitrogen molecule, N2, is triple bonded (Figure 11.8). That means that the two nitrogen atoms share 
three pairs of electrons. The first bond will be an end-to-end overlap (cr bond) and the other two bonds 
will be side-to-side overlaps (n bonds). If the end-to-end overlap are the p x orbitals, then the side-to-side 
overlaps will be the p y and p z orbitals. 

Atomic nitrogen is quite reactive but molecular nitrogen is un-reactive. The reason nitrogen molecules not 
do react readily is that the molecule is held together by an unusually strong bond - namely a triple bond. 
Approximately 78% of our atmosphere is made up of nitrogen molecules. 

337 www.ckl2.org 



£3 



:N::'N: 

Figure 11.8: Nitrogen is triple bonded. 



Lewis Formulas 



What was called "electron dot formulas" when drawing them for individual atoms become "Lewis dot for- 
mulas" or "Lewis structures" or "Lewis formulas" when drawing them for molecules. The Lewis structures 
of a molecule show how the valence electrons are arranged among the atoms of the molecule. These repre- 
sentations are named after G. N. Lewis. The rules for writing Lewis structures are based on observations 
of thousands of molecules. From experiment, chemists have learned that when a stable compound forms, 
the atoms usually have a noble gas electron configuration. Hydrogen forms stable molecules when it shares 
two electrons (sometimes called the duet rule). Other atoms involved in covalent bonding typically obey 
the octet rule. (Note: Of course, there will be exceptions.) 

Rules for Writing Lewis Structures 



Decide which atoms are bonded. 

Count all the valence electrons of all the atoms. 

Place two electrons between each pair of bonded atoms. 

Complete all the octets (or duets) of the atoms attached to the central atom. 

Place any remaining electrons on the central atom. 

If the central atom does not have an octet, look for places to form double or triple bonds. 



Example 1: 

Write the Lewis structure for water, H2O. 

Step 1: Decide which atoms are bonded. 

Begin by assuming the hydrogen atoms are bonded to the oxygen atom. i.e. Assume the oxygen atom is 
the central atom. H — O — H. 

Step 2: Count all the valence electrons of all the atoms. 

The oxygen atom has 6 valence electrons and each hydrogen has 1. The total is 8. 

Step 3: Place two electrons between each pair of bonded atoms. 

H : O : H 

Step 4: Complete all the octets or duets of the atoms attached to the central atom. 

The hydrogen atoms are attached to the central atom and hydrogen atoms require a duet of electrons and 
those duets are already present. 

Step 5: Place any remaining electrons on the central atom. 

The total number of valence electrons is 8 and we have already used 4 of them. The other 4 will fit around 
the central oxygen atom. 



www.ckl2.org 338 



h: o :h 



Is this structure correct? 

Are the total number of valence electrons correct? Yes 

Does each atom have the appropriate duet or octet of electrons? Yes 

Example 2: 

Write the Lewis structure for carbon dioxide, C02- 

Step 1: Decide which atoms are bonded. 

Begin by assuming the carbon is the central atom and that both oxygen atoms are attached to the carbon. 

Step 2: Count all the valence electrons of all the atoms. 

The oxygen atoms each have 6 valence electrons and the carbon atom has 4. The total is 16. 

Step 3: Place two electrons between each pair of bonded atoms. 

O : C : O 

Step 4: Complete all the octets or duets of the atoms attached to the central atom. 

: o : c : o: 
• # 

Step 5: Place any remaining electrons on the central atom. 

We have used all 16 of the valence electrons so there are no more to place around the central carbon atom. 

Is this structure correct? 

Is the total number of valence electrons correct? Yes 

Does each atom have the appropriate duet or octet of electrons? 

NO - each oxygen has the proper octet of electrons but the carbon atom only has 4 electrons. Therefore, 
this is NOT correct. 

Step 6: If the central atom does not have an octet, look for places to form double or triple bonds. 

Double bonds can be formed between carbon and each oxygen atom. 




Notice this time, each atom is surrounded by 8 electrons. 
Example 3: 



339 www.ckl2.org 



Write the Lewis structure for ammonia, NH3. 

Solution 

The most likely bonding for this molecule is nitrogen as the central atom and each hydrogen bonded to 
the nitrogen. Therefore, we can start by putting nitrogen in the center and put the three hydrogen atoms 
around it. 

The nitrogen atom has five valence electrons and each hydrogen atom has one so the total number of 
valence electrons is 8. 

The process for writing Lewis structures. 

h n h h:n:h h:n:h 

h ii ** ii 

The next step is to put a pair of electrons between every bonded pair of atoms so we put a pair of electrons 
between each of the hydrogen and nitrogen. The next step is to complete the octet or duet of each of 
the non-central atoms. In this case, all the non-central atoms are hydrogen and they already have a duet 
of electrons. The next step is to put all the left over electrons around the central atom. We have two 
electrons left over so they would complete the octet for nitrogen. If the central atom, at this point, does 
not have an octet of electrons, we would look for places to create a double or triple bond but in this case, 
the central atom does have an octet of electrons. The final drawing on the right is the Lewis structure for 
ammonia. 

Example 4: 

Given the skeleton structure for nitric acid, HNO3, place the electrons into a proper Lewis structure. 

Solution 

The skeleton for nitric acid has the three oxygen atoms bonded to the nitrogen and the hydrogen bonded 
to one of the oxygen atoms. The total number of valence electrons is 5 + 6 + 6 + 6+1 = 24. 

The process of writing the Lewis structure for nitric acid. 

N O H 0:N:0:H 

1 O 2 6 

:6:N:6:H : 0::N:6'.H 

3 O: 4 :0: 

•• •• 

The skeleton structure is given in "1" in the figure above. The next step is to put in a pair of electrons 
between each bonded pair - this is done in "2." So far, we have accounted for 8 of the 24 valence electrons. 
The next step is to complete the octet or duet for each of the non-central atoms. This is completed in "3." 
At that point, we have used all of the valence electrons and the central atom does not have an octet of 
electrons. The rules tell us to find a place to put a double or triple bond. For the amount of knowledge we 
have at this point, any of the three oxygen atoms is just as good as the others for a double bond, so we 
move two of the electrons around the far left oxygen atom and make a double bond between that oxygen 
and the nitrogen. Now every atom in the molecule has its appropriate octet or duet of electrons. We have 
a satisfactory Lewis structure for the nitric acid molecule. 



www.ckl2.org 340 



Lesson Summary 

• Covalent bonds are formed between atoms with relatively high electron affinity. 

• Some atoms are capable of forming double or triple bonds. 

• Multiple bonds between atoms require multiple half-filled orbitals. 

• End-to-end orbital overlaps are called sigma bonds. 

• Side-to-side orbital overlaps are called pi bonds. 

• Lewis structures are commonly used to show the valence electron arrangement in covalently bonded 
molecules. 

Review Questions 

1. Which of the following compounds would you expect to be ionically bonded and which covalently 
bonded? 

Table 11.1: 

Compound Ionic or Covalent 

CS 2 

K 2 S 

FeF% 

PF 3 

BF 3 

AIF 3 

BaS 

2. How many sigma bonds and how many pi bonds are present in a triple bond? 

3. Draw the Lewis structure for CC/4. 

4. Draw the Lewis structure for S02- 

Further Reading / Supplemental Links 

• http : //learner . org/resources/series61 . html 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. The website has one 
video that relates to this lesson called Chemical Bonds. 

Vocabulary 

covalent bond A type of bond in which electrons are shared by atoms. 

341 www.ckl2.org 



diatomic molecule A molecule containing exactly two atoms. 

double bond A bond in which two pairs of electrons are shared. 

triple bond A bond in which three pairs of electrons are shared. 

sigma bond A covalent bond in which the electron pair is shared in an area centered on a line running 
between the atoms. 

pi bond A covalent bond in which p orbitals share an electron pair occupying the space above and below 
the line joining the atoms. 

11.3 Naming Covalent Compounds 

Lesson Objectives 

• The student name covalent compounds using the IUPAC nomenclature system. 

• The student will provide formulas for covalent compounds given the IUPAC name. 

Introduction 

The systematic procedure for naming chemical compounds uses a different approach for different types of 
compounds. In a previous chapter, we have discussed the procedures for naming binary ionic compounds, 
ionic compounds involving polyatomic ions, and ionic compounds involving metals with variable oxidation 
states. In this section, we will describe the system used for covalently bonded compounds. Because of the 
large numbers of covalent compounds that may form between the same two elements, the nomenclature 
system for covalent compounds is quite different. 

The Number of Atoms in the Formulas Must be Indicated 

In naming ionic compounds, there is no need to indicate the number of atoms of each element in a 
formula because, for most cases, there is only one possible compound that can form from the ions present. 
When aluminum combined with sulfur, the only possible compound is aluminum sulfide, AI2S3. The only 
exception to this is a few variable oxidation number metals and those are handled with Roman numerals 
for the oxidation number of the metal, as in iron (II) chloride, FeCli- 

With covalent compounds, however, we have a very different situation. There are six different covalent 
compounds that can form between nitrogen and oxygen and in two of them, nitrogen has the same oxidation 
number. Therefore, the Roman numeral system will not work. Chemists devised a nomenclature system 
for covalent compounds that indicate how many atoms of each element is present in a molecule of the 
compound. 

Greek Prefixes 

In naming binary covalent compounds, four rules apply: 

1. The first element in the formula is named first using the normal name of the element. 

www.ckl2.org 342 



2. The second element is named as if it were an anion. Note: There are no ions in these compounds but 
we use the "-ide" ending on the second element as if it were an anion. 

3. Greek prefixes are used for each element to indicate the number of atoms of that element present in the 
compound. 

Table 11.2: Greek Prefixes 



Prefix 



Number Indicated 



Mono- 

Di- 

Tri- 

Tetra- 

Penta- 

Hexa- 

Hepta- 

Octa- 

Nona- 

Deca- 



1 
2 
3 
4 
5 
6 
7 
8 
9 
10 



4. The prefix "mono-" is never used for naming the first element. For example, CO is called carbon 
monoxide, not monocarbon monoxide. 



Examples 



N 2 
NO 

N0 2 

N 2 3 
N 2 A 
N 2 5 

co 2 

PaOiq 

P2S 5 



dinitrogen monoxide 
nitrogen monoxide 
nitrogen dioxide 
dinitrogen trioxide 
dinitrogen tetraoxide 
dinitrogen pentaoxide 
sulfur hexafluoride 
carbon dioxide 
tetraphosphorus decaoxide 
diphosphorus pentasulfide 



Lesson Summary 

• Covalently bonded molecules use Greek prefixes in their nomenclature. 

Review Questions 

1. Name the compound CO. 

2. Name the compound PCI3. 

3. Name the compound PC/5. 

4. Name the compound N 2 0-$. 

5. Name the compound BCI3. 

6. Name the compound SF4. 



343 



www.cki2.0rg 



7. Name the compound ChO. 

8. Write the formula for the compound sulfur trioxide. 

9. Write the formula for the compound dinitrogen tetrafluoride. 

10. Write the formula for the compound oxygen difluoride. 

11. Write the formula for the compound dinitrogen pentoxide. 

12. Write the formula for the compound sulfur hexafluoride. 

13. Write the formula for the compound tetraphosphorus decaoxide. 

Further Reading / Supplemental Links 

• http://www.mhhe.com/physsci/chemistry/animations/chang_7e_esp/bomls2_ll .swf 

• http : //www . visionlearning . com/library /module_viewer . php?mid=55 



Image Sources 



(1) CK-12 Foundation. . CC-BY-SA. 

(2) Richard Parsons. . CC-BY-SA. 

(3) Richard Parsons. . CC-BY-SA. 

(4) Richard Parsons. . CC-BY-SA. 

(5) Richard Parsons. . CC-BY-SA. 

(6) Richard Parsons. . CC-BY-SA. 

(7) CK-12 Foundation. . CC-BY-SA. 

(8) Richard Parsons. Nitrogen is triple bonded.. CC-BY-SA. 



www.ckl2.org 344 



Chapter 12 



Reactions 



12.1 Chemical Equations 

Lesson Objectives 



The student will read chemical equations and provide requested information contained in the equation 

including information about substances, reactants, products, and physical states. 

The student will convert symbolic equations into word equations and vice versa. 

The student will use the common symbols, — >,+, (s)(l)(g)(aq) appropriately. 

The student will describe the roles of subscripts and coefficients in chemical equations. 

The student will balance chemical equations with the simplest whole number coefficients. 



Introduction 

A chemical change occurs when some substances (it could even be the same substance) come into contact, 
the chemical bonds of the substances break, and the atoms that compose the compounds separate and 
re-arrange themselves into new compounds with new chemical bonds. When this process occurs, we call it 
a chemical reaction. In order to describe a chemical reaction, we need to indicate what substances were 
present at the beginning and what substances were present at the end. The substances that were present 
at the beginning are called reactants and the substances present at the end are called products. In many 
chemical reactions, it is necessary not only to name the reactants and products but also to indicate the 
phase of each substance. 

Molecules not only have a particular group of atoms arranged in an exact way, they also have a specific 
amount of potential energy associated with their chemical bonds. When reactants re-arrange the orga- 
nization of their atoms and bonds, the amount of potential energy associated with the reactant bonds is 
almost never the same as the amount of energy associated with the product bonds. Therefore, energy must 
be absorbed or given off by the reaction. For some reactions, the potential energy difference between the 
reactant bonds and product bonds must be represented in the equation. 

All of this information about a chemical reaction, reactants, products, phases, and energy changes can be 
described in words or it can be presented by a special shorthand notation devised precisely for communi- 
cating this information. 



345 www.ckl2.org 



Some Interactions Represent Chemical Reactions 

We know already that chemistry is the study of matter and that the Law of Conservation of Matter states 
that matter is conserved in a chemical reactions. But what does this really mean? When you place an ice 
cube on the counter and it melts, do you think the mass has changed? Of course you don't, but why not? 
The reason is because the mass before and after a (non-nuclear) change has to remain the same. Energy 
works the same way. The Law of Conservation of Energy states that total energy involved in a chemical 
reaction is conserved from reactants to products. 

Energy cannot be created nor destroyed in a closed system. Think of it as balancing on a seesaw (Figure 
??). If energy is absorbed by the reaction, it has to have come from somewhere in the closed system. 

Let's look at an example: 

H 2 {s) + 3312 J -» H 2 (L) 

Thus it takes 3312 Joules of energy to melt one ice cube of approximately 0.01 kg. The 3312 Joules is 
an energy unit. Where did the energy come from? The energy came from the kinetic energy in the room 
where the ice cube was melting. So the energy was absorbed by the ice cube and released by the room. 
Therefore, the total amount of energy remained constant in the room. The mass of the ice cube is the 
same as the mass of the liquid puddle. Therefore mass was conserved. See how it works? 

We will go into further depth with this concept as we explore chemical reactions in detail through the 
remainder of this unit. 




Reactants Change into Products 

Sometimes when reactants are put into a reaction vessel, a reaction will take place to produce products. 
Reactants are the starting materials, that is, whatever we have as our initial ingredients. The products are 
just that, what is produced or the result of what happens to the reactants when we put them together in 
the reaction vessel. If we think about baking chocolate chip cookies, our reactants would be flour, butter, 
sugar, vanilla, some baking soda, salt, egg, and chocolate chips. What would be the products? Cookies! 
The reaction vessel would be our mixing bowl. 

Flour + Butter + Sugar + Vanilla + Baking Soda + Eggs + Chocolate Chips — » Cookies 

Chemical reactions are similar. If sulfur dioxide is added to oxygen, sulfur trioxide is produced. Sulfur 
dioxide and oxygen (SO2 + O2) are reactants and sulfur trioxide (SO3) is the product. 

2 S0 2 ( g ) +■ 2(g ) ► 2 S0 3 |g) 



Reactants Products 

In chemical reactions, the reactants are found before the symbol "— >" and the products and found after 
the symbol "— »". The general equation for a reaction is: 

www.ckl2.org 346 



Reactants — » Products 
We will explore more about this in the next section. 

Symbol Equations are Shorthand for Word Equations 

There are a few special symbols that we need to know in order to "talk" in chemical shorthand. Remember 
when you started learning pre-algebra and you learned the order of operations. We may have all learned 
the acronym BEDMAS for brackets, exponents, division, multiplication, addition, and subtraction. The 
acronym was a shorthand method or symbolic notation for the order of operations when we want to "talk" 
in mathematical equations. We have the same thing for chemistry when we want to "talk" in chemical 
equations. In the table below is the summary of the major symbols used in chemical equations. You will 
find there are others but the main ones that we need to memorize are listed in Table 12.1. 

Table 12.1: Common Symbols 

Symbol Meaning 

— » Symbol used to separate reactants and products; 

means "to produce", can be read as "yields" 
Ex:2 H 2 + 2 -» 2 H 2 

+ Symbol used to separate reactants and/or products 

in a chemical reaction; means "is added to". 
2H 2 + 2 -» 2 H 2 
AgN0 3 + NaCl -> AgCl + NaN0 3 

(s) In the solid state Sodium in the solid state: Na(s) 

Gold in the solid state: Au(s) 

(I) In the liquid state Water in the liquid state: H20{1) 

Mercury in the liquid state: Hg(l) 

(g) In the gaseous state Helium in the gaseous state: 

He{g) 
Carbon dioxide in the gaseous state: C 02(g) 

(aq) In the aqueous state, dissolved in water to make a 

solution Sodium chloride solution: NaCl(aq) 
Hydrochloric acid solution: HCl(aq) 



Chemists have a choice of methods for describing a chemical reaction. They could draw a picture of the 
chemical reaction. 

347 www.ckl2.org 




■ ^ 



Or, they could write a word equation for the chemical reaction. 

Two molecules of hydrogen gas react with one molecule of oxygen gas to produce two molecules of water 
vapor. 

Or, they could write the equation in chemical shorthand. 

2 H 2 (g) + 2(g) -» 2 H 2 {g) 

In the symbolic equation, chemical formulas are used instead of chemical names for reactants and products 
and symbols are used to indicate the phase of each substance. It should be apparent that the chemical 
shorthand method is the quickest and clearest method for writing chemical equations. 

I could write that an aqueous solution of calcium nitrate is added to an aqueous solution of sodium 
hydroxide to produce solid calcium hydroxide and an aqueous solution of sodium nitrate. Or in shorthand 
I could write: 

Ca(N0 3 ) 2{aq) + 2 NaOH {aq) -» Ca{OH) 2{s) + 2 NaN0 3(aq) 

How much easier is that to read? Let's try it in reverse? Look at the following reaction in shorthand 
notation and write the word equation for the reaction. 

Cu (s) + AgN0 3iaci} -> Cu(N0 3 ) 2 (aq) + Ag( s ) 

The word equation for this reaction might read something like "solid copper reacts with an aqueous solution 
of silver nitrate to produce a solution of copper (II) nitrate with solid silver". 

Sample question: Transfer the following symbolic equations into word equations or word equations into 
symbolic equations. 

(a)HCl {aq) +NaOH {aq) -> NaCl (aq) + H 2 {L) 

(b) Gaseous propane (C 3 Hg) burns in oxygen gas to produce gaseous carbon dioxide and liquid water. 

(c) Hydrogen fluoride gas reacts with an aqueous solution of potassium carbonate to produce an aqueous 
solution of potassium fluoride, liquid water, and gaseous carbon dioxide. 

Solution: 

(a) An aqueous solution of hydrochloric acid reacts with an aqueous solution of sodium hydroxide to 
produce an aqueous solution of sodium chloride and liquid water. 

(b) C 3 // 8(g) + 2 (g) -> c °2( g ) + H 2 0( L ) 

(c) HF (g) + K 2 C0 3{aq) -> KF {aq) + H 2 (L) + C0 2{g) 

Lesson Summary 

• Chemical reactions can be represented by word equations or by formula equations. 

• Formula equations have reactants on the left, an arrow that is read as "yields," and the products on 
the right. 

www.ckl2.org 348 



Review Questions 

1. Mothballs are commonly used to preserve clothing in "off-season." We recognize mothballs due to 
its smell because of a chemical compound known as Naphthalene, CiqHs- What are the different 
elements found in naphthalene and how many atoms of each are found in the formula? 

2. Do you think a chemical reaction occurs every time two substances are placed together in a reaction 
vessel? 

3. Transfer the following symbolic equations into word equations. 

(a) H 2 S0 4{aq) +NaCN {aq) -> HCN {aq) + Na 2 S0 4{aq) 

(b) Cu(s) + AgN0 3{aq) -» Ag {s) +Cu(N0 3 ) 2{aq] 

(c) Fe {s) + 2 ( g ) -» Fe 2 3{s) 

4. Transfer the following equations from word equations into symbolic equations. 

(a) Solid calcium metal is placed in liquid water to produce aqueous calcium hydroxide and hydrogen 
gas. 

(b) Gaseous sodium hydroxide is mixed with gaseous chlorine to produce aqueous solutions of sodium 
chloride and sodium hypochlorite plus liquid water. 

(c) Solid xenon hexafluoride is mixed with liquid water to produce solid xenon trioxide and gaseous 
hydrogen fluoride. 

5. Did you know that you can simulate a volcanic eruption in a lab that looks like the real thing? A 
source of heat is gently placed it into a mound of ammonium dichromate. The ammonium dichromate 
decomposes to solid chromium (III) oxide, nitrogen monoxide gas, and water vapor. Write the 
symbolic reaction for the "volcanic eruption". 

Vocabulary 

reactants The starting materials in the reaction (i.e. S0 3 + H 2 — » H 2 SO^;S0 3 + H 2 = reactants). 

products The materials present at the end of a reaction (i.e. S 3 + H 2 — > H 2 SO^H 2 SO^= product). 

— » Symbol used to separate reactants and products; means to produce. Ex: 2 H 2 + 2 — » 2 H 2 0. 

+ Symbol used to separate reactants and/or products in a chemical reaction; means is added to. Examples: 
2// 2 + 2 -» 2 H 2 0;AgN0 3 + NaCl -» AgCl + NaN0 3 . 

(s) In the solid state. Examples: Sodium: Na(s); Gold: Au(s). 

(I) In the liquid state. Examples: Water: H 2 0(l); Mercury: Hg(l). 

(g) In the gaseous state. Examples: Helium: He(g); Carbon dioxide: C 02(g)- 

(aq) In the aqueous state. Examples: Sodium chloride solution: NaCl(aq). 



12.2 Balancing Equations 

Lesson Objectives 



Demonstrate the Law of Conservation of Matter in a chemical reaction. 
Explain the roles of coefficients and subscripts in a chemical reaction. 
Balance equations using the simplest whole number coefficients. 



349 www.ckl2.org 



Introduction 

Even though chemical compounds are broken up and new compounds are formed during a chemical reaction, 
atoms in the reactants do not disappear nor do new atoms appear to form the products. In chemical 
reactions, atoms are never created or destroyed. The same atoms that were present in the reactants are 
present in the products - they are merely re-organized into different arrangements. In a complete chemical 
equation, the two sides of the equation must be balanced. That is, in a complete chemical equation, the 
same number of each atom must be present on the reactants and the products sides of the 
equation. 

Subscripts and Coefficients 

There are two types of numbers that appear in chemical equations. There are subscripts which are part 
of the chemical formulas of the reactants and products and there are coefficients that are placed in front 
of the formulas to indicate how many molecules of that substance is used or produced. 

Coefficients 
Cu (s) + 2 AgN0 3 (a q ) — >Cu(N0 3 ) 2(aq) + 2 Ag (s) 




Subscripts 

The subscripts are part of the formulas and once the formulas for the reactants and products are determined, 
the subscripts may not be changed. The coefficients indicate how many molecules of each substance is 
involved in the reaction and may be changed in order to balance the equation. The equation above indicates 
that one atom of solid copper is reacting with two molecules of aqueous silver nitrate to produce one 
molecule of aqueous copper (II) nitrate and two atoms of solid silver. When you learned how to write 
formulas, it was made clear that when only one atom of an element is present, the subscript of "1" is not 
written - so that when no subscript appears for an atom in a formula, you read that as one atom. The 
same is true in writing balanced chemical equations. If only one atom or molecule is present, the coefficient 
of "1" is omitted. 

Coefficients are inserted into the chemical equation in order to balance it; that is, to make equal the total 
number of each atom on the two sides of the equation. Consider the equation representing the reaction 
that occurs when gaseous methane, C//4, is burned in air (reacted with oxygen gas), and produces gaseous 
carbon dioxide and liquid water. 



+ 



The unbalanced symbolic equation for this reaction is given below. 

Ci?4(g) + 2 (g) -* C0 2 { g ) + H 2 0( L ) Equation 1 

It is quickly apparent that equation 1 is not balanced. There are 4 hydrogen atoms in the reactants and 
only 2 hydrogen atoms in the products. The oxygen atoms in the equation are also not balanced. In order 
to balance this equation, it is necessary to insert coefficients in front of some of the substances to make the 

www.ckl2.org 350 





numbers of atoms of each element the same on the two sides of the equation. We can begin this process 
by placing a coefficient of 2 in front of the water molecule. 

11 A 






(H 



C// 4 ( g ) + 2 (g) -> C0 2 ( g ) + 2 H 2 0( L ) Equation 2 

The insertion of this coefficient balanced the hydrogen atoms (there are now 4 on each side) but the 
equation is still not completely balanced. The oxygen atoms are not balanced. There are 4 oxygen atoms 
in the products but only two in the reactants. We can now insert a coefficient of 2 in front of the oxygen 
molecule in the reactants. 




+ 




• 



C// 4 ( g ) + 2 2 ( g ) -» C0 2 ( g ) + 2 H 2 0( L ) Equation 3 

When we count up the number of each type of atom on the two sides of the equation now, we see that the 
equation is properly balanced. There is one carbon atom on each side, four hydrogen atoms on each side, 
and four oxygen atoms on each side. Equation balancing is accomplished by changing coefficients, never 
by changing subscripts. 

The Process of Balancing an Equation 

The process of writing a balanced chemical equation involves three steps. As a beginning chemistry student, 
you will not know whether or not two given reactants will react or not and even if you saw them react, 
you would not be sure what the products are without running tests to identify them. Therefore, for the 
time being, you will be told both the reactants and products in any equation you are asked to balance. 



Step 1 
Step 2 
Step 3 



Know what the reactants and products are, and write a word equation for the reaction. 
Write the formulas for all the reactants and products. 
Adjust the coefficients to balance the equation. 



You must keep in mind that there are a number elements commonly appearing in equations that, under 
normal conditions, exist as diatomic molecules (Table 12.2). 

Table 12.2: Elements that exist as diatomic molecules under normal conditions 

Element Formula for Diatomic Molecule Phase Under Normal Conditions 

Hydrogen H 2 Gaseous 

Oxygen 2 Gaseous 

Nitrogen N 2 Gaseous 

Chlorine C/2 Gaseous 

Fluorine F 2 Gaseous 

Bromine Br 2 Liquid 

Iodine I 2 Solid 

351 www.ckl2.org 



When the names of any of these elements appear in word equations, you must remember that the name 
refers to the diatomic molecule and insert the diatomic formula into the symbolic equation. If, under 
unusual circumstances, it was desired to refer to the individual atoms of these elements, the text would 
refer specifically to atomic hydrogen or atomic oxygen and so on. 

Sample Problem 1 

Write a balanced equation for the reaction that occurs between chlorine gas and aqueous sodium bromide 
to produce liquid bromine and aqueous sodium chloride. 

Step 1: Write the word equation (keeping in mind that chlorine and bromine refer to the diatomic 
molecules). 

Chlorine + sodium bromide — » bromine + sodium chloride 

Step 2: Substitute the correct formulas into the equation. 

Cl 2 + NaBr -> Br 2 + NaCl 

Step 3: Insert coefficients where necessary to balance the equation. 

By placing a coefficient of 2 in front of the NaBr, we can balance the bromine atoms and by placing a 
coefficient of 2 in front of the NaCl, we can balance the chloride atoms. 

Cl 2 + 2 NaBr -> Br 2 + 2 NaCl 

A final check (always do this) shows that we have the same number of each atom on the two sides of the 
equation and we do not have a multiple set of coefficients so this equation is properly balanced. 

Sample Problem 2 

Write a balanced equation for the reaction between aluminum sulfate and calcium bromide to produce 
aluminum bromide and calcium sulfate. It might be worthwhile to note here that the reason you memorized 
the names and formulas for the polyatomic ions is that these ions usually remain as a unit throughout 
many chemical reactions. 

Step 1: Write the word equation. 

Aluminum sulfate + calcium bromide — » aluminum bromide + calcium sulfate 

Step 2: Replace the names of the substances in the word equation with formulas. 

Al 2 (S A ) 3 + CaBr 2 -» AlBr 3 + CaS A Equation 1 

Step 3: Insert coefficients to balance the equation. 

In order to balance the aluminum atoms, we must insert a coefficient of 2 in front of the aluminum 
compound in the products. 

Al 2 (S 04)3 + CaBr 2 -> 2 AlBr 3 + CaS 4 Equation 2 

In order to balance the sulfate ions, we must insert a coefficient of 3 in front of the CaSO^ in the products. 
Al 2 (S0 4 ) 3 + CaBr 2 -» 2 AlBr 3 + 3 CaSO± Equation 3 

www.ckl2.org 352 



In order to balance the bromine atoms, we must insert a coefficient of 3 in front of the CaBr2 in the 
reactants. 

Ah{S0 4 )3 + 3 CaBr 2 -> 2 AlBr 3 + 3 CaS0 4 Equation 4 

The insertion of the 3 in front of the CaBr2 in the reactants also balances the calcium atoms in the CaS O4 
in the products. A final check shows 2 aluminum atoms on each side, 3 sulfur atoms on each side, 12 oxygen 
atoms on each side, 3 calcium atoms on each side, and 6 bromine atoms on each side. This equation is 
balanced. 

Sample Problem 3 

Balance the following skeletal equation. (The term "skeletal equation" refers to an equation that has the 
correct formulas but has not yet had the proper coefficients added.) 

Fe(N0 3 ) 3 + NaOH -> Fe(OH) 3 + NaN0 3 (skeletal equation) 

We can balance the hydroxide ion by inserting a coefficient of 3 in front of the NaOH on the reactant side. 
Fe{N0 3 ) 3 + 3 NaOH -> Fe(OH) 3 + NaN0 3 (intermediate equation) 

Then we can balance the nitrate ions by inserting a coefficient of 3 in front of the sodium nitrate on the 
product side. 



Counting the number of each type of atom on the two sides of the equation will now show that this equation 
is balanced. 

Simplest Whole Number Coefficients 

Chemical equations should be balanced with the simplest whole number coefficients that balance the 
equation. Here is the properly balanced equation from the previous section. 

AZ 2 (S 6*4)3 + 3 CaBr 2 -» 2 AlBr 3 + 3 CaSOt Equation 4 

You should note that the equation in the previous section would have the same number of atoms of each 
type on each side of the equation with the following set of coefficients. 

2 Al 2 {S0 4 ) 3 + 6 CaBr 2 -» 4 AlBr 3 + 6 CaS0 4 Equation 5 

You should count the number of each type of atom on each side of the equation to confirm that this 
equation is "balanced". While this set of coefficients do "balance" the equation, they are not the lowest 
set of coefficients possible that balance the equation. We could divide each of the coefficients in this 
equation by 2 and get a another set of coefficients that are whole numbers and also balance the equation. 
Since it is required that an equation be balanced with the lowest whole number coefficients, equation 5 is 
NOT properly balanced. The properly balanced equation for this reaction is equation 4. When you have 
finished balancing an equation, you should not only check to make sure it is balanced, you should also 
check to make sure that it is balanced with the simplest set of whole number coefficients possible. 

353 www.ckl2.org 



Let's consider another equation for balancing; the combustion of octane, C 8 His- 

C 8 H 18 + 2 ^> C0 2 + H 2 Equation 1 

To begin balancing this equation, we note that there are 8 carbon atoms on the reactant side and only 1 
carbon on the product side. We can balance the carbon atoms by inserting a coefficient of 8 in front of the 
carbon dioxide on the product side. 

C 8 H i8 + <9 2 -> 8 C0 2 + H 2 Equation 2 

Next, we note that there are 18 hydrogen atoms in the reactants and only 2 hydrogen atoms in the products. 
We can balance the hydrogen atoms by inserting a coefficient of 9 in front of the water molecule on the 
product side. 

C 8 H 18 + 2 -> 8 C0 2 + 9 H 2 Equation 3 

Now when we count up the number of oxygen atoms in the products, we find there are 25 oxygen atoms. 
What coefficient can you insert in front of the 2 molecule in the reactants to have 25 oxygen atoms? Since 
the oxygen molecule has a subscript of 2, the only coefficient for this molecule that will produce 25 oxygen 
atoms is a coefficient of y. If we insert y as the coefficient for the oxygen molecule, we get a balanced 
equation but we do not get an equation that is balanced with the simplest whole number coefficients. 

25 
C 8 H 18 + — 2 -> 8 C0 2 + 9 H 2 Equation 4 

In order to have whole number coefficients for this equation, we must go through the equation and multiply 
all the coefficients by 2. That produces the coefficients shown in equation 5. 

2 C 8 H 18 + 25 2 -» 16 C0 2 + 18 H 2 Equation 5 

Now the equation is balanced with the simplest whole number coefficients and is a properly balanced 
equation. 

Sample Problem 

Given the following skeletal (un-balanced) equations, balance them. 

(a) CaC0 3 ^ — > CaOi s \ + C0 2 r g \ 

(b) H 2 SO i{aq) +Al{OH) 3(aq) ^Al 2 {SO A ) 3{aq) +H 2 {L) 

(c) Ba(N0 3 ) 2 ( aq ) + Na 2 C0 3{aq) -> BaC0 3{aq) + NaN0 3{aq) 

(d) C 2 // 6(g ) + 2 (g) -» C0 2 ( S ) + H 2 0( L ) 
Solutions 

(a) CaC0 3 ( s ) — » CaOi s \ + C0 2 t g \ (sometimes the equation balances with all coefficients of 1) 

(b) 3 H 2 SO A(aq) + 2 Al{OH) 3{aq) -> Al 2 (S0 4 ) 3{aq) + 6 H 2 {L) 

(c) Ba(N0 3 ) 2[aq) + Na 2 C0 3{aq) -» BaC0 3(aq) + 2 NaN0 3(aq) 

(d) 2 C 2 H 6{g) + 7 2{g) -» 4 C0 2(g) + 6 H 2 {L) 

www.ckl2.org 354 



The Conservation of Mass in Chemical Reactions 

Both the numbers of each type of atom and the mass are conserved during chemical reactions. An exam- 
ination of a properly balanced equation will demonstrate that mass is conserved. Consider the following 
reaction. 

Fe(N0 3 ) 3 + 3 NaOH -> Fe{OH) 3 + 3 NaN0 3 

You should demonstrate that this equation is balanced by counting the number of each type of atom on 
each side of the equation. 

We can also demonstrate that mass is conserved in this reaction by determining the total mass on the two 
sides of the equation. 

Fe{N0 3 ) 3 + 3 NaOH -^ Fe(OH) 3 + 3 NaN0 3 

Reactant Side Mass 

1 molecule of Fe{N0 3 ) 3 x molecular weight = (1) (241.9 daltons) = 241.9 daltons 

3 molecules of NaOH x molecular weight = (3) (40.0 daltons) = 120. daltons 

Total mass for reactants = 241.9 daltons + 120. daltons = 361.9 daltons 

Product Side Mass 

1 molecule of Fe{OH) 3 x molecular weight = (1)( 106.9 daltons) = 106.9 daltons 

3 molecules of NaN0 3 x molecular weight = (3) (85.0 daltons) = 255 daltons 

Total mass for products = 106.9 daltons + 255 daltons = 361.9 daltons. 

As you can see, both number of atom types and mass are conserved during chemical reactions. A group of 
20 objects stacked in different ways will still have the same total mass no matter how you stack them. 

Lesson Summary 

• To be useful, chemical equations must always be balanced. 

• Balanced chemical equations must have the same number and type of each atom on both sides of the 
equation. 

• The coefficients in a balanced equation must be the simplest whole number ratio. 

• Mass is always conserved in chemical reactions. 

Review Questions 

1. Explain in your own words why it is essential that subscripts remain constant but coefficients can 
change. 

2. Which set of coefficients will properly balanceLthe following^ equation? 

(a) 1,1,1,1 

(b) 1,3,2,2 

(c) 1,3.5,2,3 

(d) 2,7,4,6 

3. When properly balanced, what is the, sum oXall thqcoefficients^in the following chemical equation? 

355 www.ckl2.org 



(a) 4 

(b) 7 

(c) 9 

(d) None of the above 

4. When the following equation is balanced, what is the coefficient found in front of the 2 ? 

Pi + 2 + H 2 -» H 3 POi 

(a) 1 

(b) 3 

(c) 5 

(d) 7 

5. Balance the following equations. 

(a) XeF 6(s) + H 2 {L) -» XeO s(s) + HF (g) 

(b) C« w + AgN0 3(aq) -> Ag {s) + CM(W0 3 )2( a? ) 

(c) Fe is) + 2 ( g ) -» Fe 2 3{s) 

(d) A7(0#) 3 + Mg 3 ( J P0 4 )2 -» A/P0 4 + Mg{OH) 2 

Further Reading / Supplemental Links 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 8-1 is on Balancing Equations. 

• http : //www . f ordhamprep . org/gcurran/sho/sho/lessons/lesson81 . htm 

Vocabulary 

law of conservation of matter Matter is neither created nor destroyed in chemical reactions. 

skeletal equation a chemical equation before it has been balanced 

balanced chemical equation a chemical equation in which the number of each type of atom is equal 
on the two sides of the equation 



12.3 Types of Reactions 

Lesson Objectives 



• Identify the types of reactions. 

• Predict the products in different types of reactions. 

• Distinguish between the different types of reactions. 

• Write balanced chemical equations and identify the reaction type given only the reactants. 

www.ckl2.org 356 



Introduction 

Chemical reactions are classified into groups to help us analyze them and also to help us predict what the 
products of the reaction will be. The four major types of chemical reactions are synthesis, decomposition, 
single replacement, and double displacement. Sometimes other names are used for these basic types of 
reactions but the same four are always listed. There are also some sub-groups under these four but we will 
concentrate on the basic four groups. 

Synthesis Reactions 

A synthesis reaction is one in which two or more reactants combine to make ONE type of product. 

General equation: A + B — » AB 

Synthesis reactions occur as a result of two or more simpler elements or molecules combining to form a more 
complex molecule. Look at the example below. Here two elements (hydrogen and oxygen) are combining 
to form one product (water). 

Example: 2 H 2 (g) + 2 { g ) -> 2 H 2 0( L ) + energy 

(notice that energy is a product and therefore this synthesis reaction is exothermic) 

We can always identify a synthesis reaction because there is only one product or one substance present 
after the arrow (on the right hand side) of the reaction. 

We can write the synthesis reaction for sodium chloride just by knowing the elements that are present in 
the product. 

2 Na(s) + Cl 2 (g) -> 2 NaCl(s) 

You will recognize synthesis reactions by the fact that there is only one product. You should be able to 
write a formation synthesis reaction if you are given a product by picking out its elements and writing the 
reaction (just like we did with NaCl). Also, if you are given elemental reactants and told that the reaction 
is a synthesis reactant, you should be able to predict the products. 

Sample Question: 

(a) Write a synthesis reaction to produce silver bromide, AgBr. 

(b) Predict the products for the following reaction: C0 2 ^ + H 2 Ot L \ — » 

(c) Predict the products for the following reaction: Li 2 0( s \ + C0 2 ( g ) — > 
Solution: 

(a) 2 Ag (s) + Br 2{L) -» 2 AgBr (s) 

(b) C0 2{g) + H 2 {L) -» H 2 C0 3H) 

(c) Li 2 {s) + C0 2{g) -> Li 2 C0 3 ( s) 

Decomposition Reactions 

The best way to remember a decomposition reaction is that for all reactions of this type, there is only one 
reactant. When ONE type of reactant breaks down to form two or more products, we have a decomposition 
reaction. 

General Equation: AB — > A + B 

357 www.ckl2.org 



Look at the example below for the decomposition of ammonium nitrate to dinitrogen oxide and water. 

Example: energy + NH A NO^ s ) -» N 2 0( g ) + 2 H 2 0( g ) 

(Notice the energy is a reactant and therefore this decomposition reaction is endothermic) 

Notice the one reactant jNH/jNO^a, is on the left of the arrow and there is more than one on the right side 
of the arrow. This is the exact opposite of the synthesis reaction type. 

When studying decomposition reactions, we can predict reactants in a similar manner as we did for synthesis 
reactions. Look at the formula for magnesium nitride, Mg 3 N 2 . What elements do you see in this formula? 
You see magnesium, Mg, and nitrogen, N 2 . Now we can write a decomposition reaction for magnesium 
nitride. 



Mg 3 N 2{s) -> 3 Mg {s) + N 2 ( g ) 



Notice there is only one reactant. 
Sample Question: 

(a) Write a decomposition reaction for aluminum oxide, Al 2 3 . 

(b) Predict the products for the following reaction: Ag 2 S — > 

(c) Predict the products for the following reaction: MgO — » 
Solution: 

(a) 2 Al 2 3 -»4 Al + 3 2 

(b) Ag 2 S -^2Ag + S 

(c) 2 MgO -> 2 Mg + 2 

Single Replacement (metal) 

A third type of reaction is the single replacement reaction. In single replacement reactions one element 
reacts with one compound to form products. The single element is said to replace an element in the 
compound when products form, hence the name single replacement. 

There are actually three different types of single replacement reactions; 1) the single element is a metal and 
replaces the metal of the compound in the second reactant, 2) the single element is a metal and replaces 
the hydrogen of the compound in the second reactant which is always an acid, and 3) the single element 
is a non-metal and replaces the non-metal in the compound. The single replacement reaction in which 
hydrogen is replaced in an acid will be covered in detail in a later section. In this section, let us focus 
on single replacement reactions where the element reactant replaces the metal in the compound reactant. 
Look at the general reaction below. 

General equation: A + BC — » B + AC 

(element replaces cation of compound) 

Example: Zn {s) + Cu(N0 3 ) 2 ( aq ) -» Zn{N0 3 ) 2 ( aq) + Ci4( s ) 

Notice there is only one reactant that is an element and one reactant that is a compound. 

When studying single replacement reactions, we can predict reactants in a similar manner as we did for 
synthesis and decomposition reactions. Remember the oxidation numbers learned previously. This prior 
knowledge will become especially important here when predicting products for the rest of this chapter. 
Let's try an example where we will predict a series of products for an element reacting with a compound 
where the element will replace the metal ion. 

www.ckl2.org 358 



Suppose that we know a single replacement reaction will occur between solid aluminum and solid iron (III) 
oxide. 

A/ w + Fe 2 Hs) -> 

In order to predict the products we need to know that aluminum will replace iron and form aluminum 
oxide. Aluminum has a charge of +3 (it is in group 3A) and oxygen has a charge of -2 (it is in group 6A). 
The compound formed between aluminum and oxygen, therefore, will be Al-zO^uy Since iron is replaced 
in the compound by aluminum, the iron will now be the single element in the products. The unbalanced 
equation will be: 

Al {s) + Fe 2 3{s) -> Al 2 3(s) + Fe {s) 

and the balanced equation will be: 

2 A/ w + Fe 2 3{s) -» Al 2 3{s) + 2 Fe (s) 

Sample Question: 

(a) Write a single replacement reaction for the reaction between zinc solid and lead (II) nitrate solution to 
produce zinc nitrate solution and solid lead. 

(b) Predict the products for the following reaction: Fe + CuS O4 — > 
(b) Predict the products for the following reaction: AI + CuCl 2 — > 
Solution: 

(a) Zn {s) + Pb(N0 3 ) 2(aq) -> Pb (s) + Zn(N0 3 ) 2(aq) 

(b) Fe {s) + CuSO i{aq) -> FeS0 4{aq) + Cu {s) 

(c) 2 Al + 3 CuCl 2 -» 3 Cm + 2 A/C/3 

Single Replacement (many metals with acid) 

These reactions are the same as those studied in the last section except the compound in the reactant 
side of the equation is always an acid. Since you may not have studied acids yet, you should consider an 
acid to be a compound in which hydrogen is combined with a anion. Therefore, in this section, the single 
replacement reactions will be where the element reactant replaces the hydrogen in the compound reactant. 
Look at the general reaction below. 

General equation: A + 2 HC — » AC 2 + H 2 (element replaces hydrogen of compound) 

Example: Zn {s) + 2 HBr {aq) -» ZnCl 2{aq) + # 2 (g) 

When studying these single replacement reactions, we can predict reactants in a similar manner as we did 
for the other types of single replacement reactions. Look at the reaction below. Since HCl is a compound 
that has hydrogen combined with an anion, it is an acid. 

Mg {s) + 2 HCl (aq) -» 

In order to predict the products, we need to know that magnesium will replace hydrogen and form mag- 
nesium chloride. Magnesium has a charge of +2 (it is in group 2) and chlorine has a charge of -1 (it 
is in group 7A). Therefore, the compound formed will be MgCl 2 . When hydrogen is replaced from the 

359 www.ckl2.org 



compound, it will appear in the products as elemental hydrogen, H 2 . All we need to do after the products 
are determined is to balance the equation. 

Mg(s) + 2 HCl(aq) -> MgCl 2 (s) + H 2 (g) 

Notice there is one reactant that is an element (Mg) and one reactant that is an acid compound. The Mg 
has replaced the hydrogen in the HCl in the same manner as the Zn replaced the hydrogen in the HBr in 
the example above. 

Sample Question: 

(a) Write a single replacement reaction for the reaction between iron solid and hydrochloric acid solution 
to produce iron(II) chloride solution and hydrogen gas. 

(b) Predict the products for the following reaction: Znt s \ + H 2 S 0^ aq ) — » 

(c) Predict the products for the following reaction: Ali s \ + HNO^ aq ^ — » 
Solution: 

(a)Fe w + 2 HCl {aq) -> FeCl 2{aq) + H 2{g) 
(b)Zn( s ) + H 2 S 0^ aq ) -» ZnS 0^ aq) + H 2 ( g ) 
(c)2A/ w + 6 HN0 3{aq) -» 2 A/(iV0 3 ) 3 W) + 3 H 2{g) 

Single Replacement (non-metal) 

In this section, we will focus on the final type of single replacement reactions where the element reactant 
replaces the non-metal (or the anion) in the compound reactant. Look at the general reaction below. 

General equation: A + BC — » C + BA (element replaces anion of compound) 

Example: Cl 2 ( g ) + 2 Kl^ aq) -> 2 KCl {aq) + l 2 ( s ) 

Notice, in the example, the chlorine replaced the iodine to produce solid iodine as a product. We can predict 
the products for these single replacement reactions in a similar manner as for all other single replacement 
reactions. The only difference here is that we have to remember that we are replacing the anion of the 
compound rather than the cation. Look at the reaction below between chlorine gas and sodium bromide. 
This is an actual extraction method for extracting bromine from the ocean water found to contain sodium 
bromide. Can you complete the reaction? 

Cl 2{g) + NaBr {aq) -> 

In order to predict the products of this reaction we need to know that chlorine will replace bromine and 
form sodium chloride. Sodium has a charge of +1 (it is in group 1) and chlorine has a charge of -1 (it is 
in group 17). The compound formed will be NaCl. 

Cl 2 ( g ) + 2 NaBr {aq) -> 2 NaCl {aq) + Br 2{L) 

Notice as with all of the other single replacement reactions, the reactants include one element and one 
compound and the products contain one element and one compound. This is the determining factor for 
identifying whether you have a single replacement reaction. 

Sample Question: 

(a) Write a single replacement reaction for the reaction between sodium iodide solution and bromine 
solution to produce sodium bromide solution and solid iodide. 

www.ckl2.org 360 



(b) Predict the products for the following reaction: Br 2 ( aq ) + Kh aq \ — > 
(b) Predict the products for the following reaction: Mgl 2 ^ + Cl^ig) — » 
Solution: 

(a) 2 Nal {aq) + Br 2{L) -> 2 NaBr (aq) + / 2(s) 

(b) 5r 2(a(? ) + 2 X7 (flg ) -> 2 X"fir (a?) + / 2( ,) 

(c) Mgl 2(aq) + C/ 2 ( g ) -> MgCl 2 ( aq ) + h(s) 

Double Replacement 

For double replacement reactions two reactants will react by having the cations exchange places with the 
anions. The key to this type of reaction, as far as identifying it over the other types, is that it has two 
compounds as reactants (or before the arrow). 

This type of reaction is more common than any of the others and there are many different types of double 
replacement reactions. Some double replacement reactions are more common than others. For example, 
precipitation and neutralization reactions are two of the most common double replacement reactions. 
Precipitation reactions are ones where two aqueous compound reactants combine to form products where 
one of the products is an insoluble solid. A neutralization reaction is one where the two reactant compounds 
are an acid and a base and the two products are a salt and water (i.e. acid + base -» salt + water). 

General equation for a double displacement reaction: AB + XY — ^ XB + AY 

t f 

In this reaction, the reactant compounds exchanged partners. A broke up with B and joined Y and X broke up with Y and 
joined B. 

Example: AgN0 3{aq) + NaCl {aq) -^ AgCl {s) + NaN0 3(aq) 
(this is a precipitation reaction because AgCL s \ is formed) 

2 NaOH {aq) + H 2 SO A{aq) -» Na 2 S0 4{aq) + 2 H 2 {L) 

(this is a neutralization reaction because the acid, H 2 S04, is neutralized by the base, NaOH) 

In order to write the products for a double displacement reaction, you must be able to determine the 
correct formulas for the new compounds. Let's practice with an example or two. 

A common laboratory experiment involves the reaction between lead(II) nitrate and sodium iodide, both 
clear solutions. Here is the start of the reaction. 

Pb(N0 3 ) 2{aq} + Nal {aq) -> 

Now, predict the products from your knowledge of oxidation numbers (or charges). 

We know that the cations exchange anions. We now have to look at the charges of each of the cations and 
anions to see what the products will be. 

We should presume the oxidation number of the lead will remain 2+ and since iodine forms ions with an 
oxidation number of 1-, one product will be Pbl 2 . The other product will form between sodium ions whose 
oxidation number is 1+ and nitrate ions whose oxidation number is 1-. Therefore, the second product will 
be NaN0 3 . Once the products are written in, the equation can be balanced. 

361 www.ckl2.org 



Pb(N0 3 )2(aq) + 2 Nal {aq) -» Pbl 2{s) + 2 NaN0 3{aq) . 

The experiment produces a brilliant yellow precipitate. If you have use of a solubility table, it is easy to 
determine that the precipitate will be the lead (II) iodide. Lead compounds tend to precipitate and sodium 
compounds are always soluble so we would know that the Pbl 2 is the brilliant yellow precipitate. 

Look at the reaction between acetic acid and barium hydroxide below. 

HC 2 H 3 2{ag) + Ba{OH) 2{aq) -» ? 

Predict the products in the same manner as you did in the previous reaction by having the cations exchange 
places and writing the correct formulas for the products formed. 

HC 2 H 3 2 (aq) + Ba(OH) 2 ( aq ) -> Ba{C 2 H 3 2 ) 2{aq) + H 2 0^ (not balanced) 

Therefore the final reaction will be: 2 HC 2 H 3 2 ^ + Ba(OH) 2 ^ aq -j — > Ba(C 2 H 3 2 ) 2 ^ + 2 H 2 0^ (balanced) 

This is an acid-base reaction yielding the salt, barium acetate, and water. Notice that HOH and H 2 are 
the same. 

Sample Question: 

(a) Write a double replacement reaction for the reaction between calcium chloride solution and potassium 
hydroxide solution to produce potassium chloride solution and a precipitate of calcium hydroxide. 

(b) Predict the products for the following reaction: AgN0 3 ( aq ^ + NaCh aq \ -» 

(c) Predict the products for the following reaction:,FeC/3(ag) + KOH(aq) — > 
Solution: 

(a) CaCl 2{aq) + 2 KOH (aq) -» Ca(OH) 2{s) + 2 KCl {aq) 

(b) AgN0 3(aq) + NaCl {aq) -» AgCl {s) + NaN0 3(aq) 

(c) FeCl 3 (aq) + 3 KOH{aq) -» Fe{OH) 3{s) + 3 ^C/ (a?) 

Combustion Reactions 

A special type of single displacement reaction deserves some attention. These reactions are combustion 
reactions, specifically ones that involve fuels such as hydrocarbon compounds. In a particular branch of 
chemistry, known as organic chemistry, we study compounds known as hydrocarbons. You might remember 
this from previous science classes. In any event, hydrocarbons (hydrogen + carbon) represent the major 
components of all organic material including fuels. Combustion reactions usually have the same products, 
C0 2 and H 2 0, and one of its reactants is always oxygen. In other words, the only part that changes from 
one combustion reaction to the next is the actual hydrocarbon that combusts. The general equation is 
given below. Notice the oxygen, carbon dioxide, and water parts of the reaction are listed for you to show 
you how these reactants and products remain the same from combustion reaction to combustion reaction. 

General equation: Hydrocarbon + 02(g) — * C0 2 ^ + H 2 0^ 

Also, combustion reactions are all exothermic. Small hydrocarbons (hydrocarbons having 1 to 4 carbon 
atoms in them) are gases at room temperature; all of the rest are liquids. Look at the reaction for the 
combustion of octane, CsH±s, below. Octane has 8 carbon atoms hence the prefix "oct". 

Example: 2 C 8 #i 8{L) + 25 2(g) -> 16 C0 2{g) + 18 H 2 {L) 

Now you try one. Write the combustion reaction for the combustion of hexane, Ce^i4(L)- 

www.ckl2.org 362 



Step 1: We know the general equation: Hydrocarbon + 02(g) ~ * ^02(g) + H 2 On\, now put in our hydro- 
carbon, hexane, into the general equation. 

C&H\A{L) + 02(g) -» C0 2 ( g ) + #20(L) 

Step 2: We now have to balance the reaction. Remember we have to start with the carbon atoms, then 
the hydrogen and finally oxygen. Do you know why? Carbon is first because it is considered a main 
element - so we put a coefficient of 6 in front of the carbon dioxide in the products. Then if we look 
at hydrogen and oxygen, hydrogen is in 2 compounds (species: CqHu and H 2 0) and oxygen is in 3 
elements/compounds (species: 2 ,C0 2 , and H 2 0). Since hydrogen appears fewer times, we do hydrogen 
before oxygen. Therefore, we would put a coefficient of 7 in front of the water in the products. 

C6#i4(L) + 02(g) -» 6 C0 2 ( g ) + 7 H 2 0( L ) 

In order to balance the oxygen, we have to insert a coefficient of -j in front of the oxygen gas in the 
reactants. 

19 

C6#14(L) + Y°2fe) -* 6 C02(g) + ' #20(L) 

Step 3: Remove the fraction. (Multiply the entire reaction by 2) 

2 C 6 #i4(L) + 19 2{g) -» 12 C0 2{g) + 14 H 2 (L) 

Sample Question: Write the combustion reaction for the combustion of butane, C&H\q. 
Solution: 

2 C^HiQ^g) + 13 2 (g) -* 8 C0 2 ( g ) + 10 //20(l) 

The reactions we have done so far are what are referred to as complete combustion. Complete combustion 
reactions occur when there is enough oxygen to burn the entire hydrocarbon. This is why there are only 
carbon dioxide and water as products. Have you ever, though, been in a lab, heating a beaker and seen the 
black soot appear on the bottom of the beaker. Or seen the black puffs of smoke come out from the exhaust 
pipe of a car? If there is not enough oxygen the result is an incomplete combustion reaction with COt g \ 
and C soot also formed as products. Incomplete combustion reactions are actually quite dangerous because 
one of the products in the reaction is not carbon dioxide but carbon monoxide. Carbon monoxide is the 
gas that causes people to become sleepy when they breathe it and then eventually, when the concentration 
in the blood becomes too high, the person dies. 

Sample Question: Identify whether each of the following reactions are complete or incomplete combustions 
and then balance each reaction. 

(a) C 7 #i 6 ( L ) + 2 (g) -» C0 2 ( g ) + H 2 0( L ) 

(b) C 3 // 8(g ) + 02( g ) -» C0 2 ( g ) + #20(Z,) 

(c) CH A(g) + 2{g) -> CO {g) (g) + H 2 {L) 

(d) C 5 // 12 (L) + 02(g) -> C 02(g) + #20(L) 

(e) C 2 # 6 ( g ) + 2 (g) -» CO^-j + #20(L) 

Solution: 

(a) Complete; C 7 H 16 ^ + 11 2(g) -» 7 C0 2 ( g ) + 8 # 2 0(l) 

363 www.ckl2.org 



(b) Complete; C 3 H 8{g) + 5 2{g) -> 3 C0 2{g) + 4 H 2 {L) 

(c) Incomplete; 2 CH 4(g) + 3 2fe) -» 2 C0 fe) (g) + 4 # 2 (L) 

(d) Complete; C 5 H 12 ( L ) + 8 2 ( g ) -» 5 C0 2 (s) + 6 H 2 0( L ) 

(e) Incomplete; 2 C 2 H 6 ^ + 5 2 (g) ~* 4 C#(g) + 6 # 2 0(l) 

Lesson Summary 

• There are five types of chemical reactions (Table 12.3): 

Table 12.3: 



Reaction Name Reaction Description 

Synthesis: two or more reactants make one product. 

Decomposition: one type of reactant makes two or more products. 

Single replacement: one element species reacts with one compound 

species to form products. 

Double replacement: two compounds act as reactants. 

Combustion: a fuel reactant reacts with oxygen gas. 



Review Questions 

1. When balancing combustion reactions, did you notice a consistency relating to whether the number 
of carbons in the hydrocarbon was odd or even? 

2. Distinguish between synthesis and decomposition reactions. 

3. When dodecane, C\qH 22 , burns in excess oxygen, the products would be 

(a) C0 2 + 2H 2 

(b) CO + H 2 

(c) C0 2 + H 2 

(d) c// 4 o 2 

4. In the decomposition of antimony trichloride, which of the following products and quantities will be 
found? 

(a) An + Cl 2 

(b) 2 An + 3 Cl 2 

(c) Sb + Cl 2 

(d) 2 Sb + 3 Cl 2 

5. Acetylsalicylic acid (Aspirin ), CgHgO^s), is produced by reacting acetic anhydride, C^HqO^I), 
with salicylic acid, CjHqO^s). The other product in the reaction is water. Write the balanced 
chemical equation. 

6. When iron rods are placed in liquid water, a reaction occurs. Hydrogen gas evolves from the container 
and iron(III) oxide forms onto the iron rod. 

(a) Write a balanced chemical equation for the reaction. 

(b) What type of reaction is this? 

7. A specific fertilizer is being made at an industrial plant nearby. The fertilizer is called a triple 
superphosphate and has a formula Ca{H 2 PO '4)2- It is made by sand and clay that contains phosphate 

www.ckl2.org 364 



and then treating it with a calcium phosphate solution and phosphoric acid. The simplified reaction is 
calcium phosphate reacting with the phosphoric acid to yield the superphosphate. Write the balanced 
chemical reaction and name the type. 



Further Reading / Supplemental Links 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 8-2 is on the Classification of Chemical Reactions, http : //www . f ordhamprep . org/gcurran/ 
sho/sho/lessons/lesson82 . htm 

• http : //library .thinkquest .org/2923/react .html 

• http://en.wikipedia.org 



Vocabulary 

synthesis A synthesis reaction is one in which two or more reactants combine to make ONE type of 
product. (A + B ^ C). 

decomposition A decomposition reaction is one in which ONE type of reactant breaks down to form 
two or more products. (C — » A + B). 

single replacement (metal) In a single replacement (metal) reaction, one element replaces the metal 
cation of the compound reactant to form products. {A + BC — » AC + B). 

single replacement (many metals with acid) In a single replacement (many metals with acid) reac- 
tion, one element replaces the hydrogen cation of the compound (which is an acid) reactant to form 
products. Example: A + 2 HC — > AC2 + Hi- 

single replacement (non-metal) In a single replacement (non-metal) reaction, one element replaces 
the non-metal (anion) of the compound reactant to form products. (XY + Z — » XL + Y). 

double replacement For double replacement reactions two reactants will react by having the cations 
replace the anions. (AB + XY -» AY + XB). 

combustion (complete) Combustion is the burning in oxygen, usually a hydrocarbon. ( fuel + O2 — » 
CO2 + H2O). 

combustion (incomplete) Incomplete combustion is the inefficient burning in oxygen, usually a hydro- 
carbon. Inefficient burning means there in not enough oxygen to burn all of the hydrocarbon present, 
sometimes carbon (soot) is also a side product of these reactions. ( fuel + O2 — » CO2 + H2O). 

hydrocarbons Compounds containing hydrogen and carbon. 

365 www.ckl2.org 



Labs and Demonstrations for Chemical Reactions 
Chemical Reactions in Microscale 

Investigation and Experimentation Objectives 




In this activity, the student will see the various types of chemical reactions and will associate particular 
observations with specific reaction types. 

Purpose 

To write balanced chemical reactions, identifying the type of reaction, and the physical characteristics that 
indicate a reaction has taken place. 

Background 

Within the five main types of chemical reactions studied in the Chemical Reactions chapter, four of these 
(synthesis, decomposition, single replacement, and double replacement), have subgroups of reactions that 
can be classified. These subgroups are known as precipitation reactions, neutralization reactions, combus- 
tion reactions, and the like. Physical changes such as: the formation of a precipitate (hence the precipitation 
reaction subclass), change in color, gas formation, change in temperature, tell us that a reaction has taken 
place. As well, within each of these reactions, whether an observable physical change has occurred or not, 
the Law of Conservation of mass is always maintained. 

In this lab you are given seven nitrate solutions with which you are going to react seven sodium and one 
ammonium solutions. Planning is everything! Added to this you are only working with 10 drops of solution 
in total. 

Materials 

0.1 mol/L Cu{N0 3 ) 2 

0.1 mol/L Na 2 C0 3 

0.1 mol/L Pb(N0 3 ) 2 

0.1 mol/L Na 2 S0 4 

0.1 mol/L Ni{N0 3 ) 2 

0.1 mol/L NaCl 

0.1 mol/L Co(N0 3 ) 3 

0.1 mol/L Nal 

0.1 mol/L Fe(N0 3 ) 3 

0.1 mol/L Na 2 CrC>4 

0.1 mol/L AgN0 3 

0.1 mol/L Na 2 Cr 2 7 

0.1 mol/L HN0 3 

0.1 mol/L NaOH 

0.1 mol/L NH^OH 

H 2 

24-well micro plate 

toothpicks 

beral pipettes 

24-well micro plate (x 4) 



www.ckl2.org 366 



Table 12.4: Data Table 



A 
B 
C 
D 
E 
F 
G 



Safety 

Avoid contact with the solutions. If solutions get on your skin, rinse the area thoroughly with running 
water. 

Procedure 

Part 1: Place 4 micro plates in a grid so that you have at least 8 wells in a line and 7 lines down. 

Part 2: In wells Al through to A8, add 5 drops of 0.1 mol/L Cu(N0 3 ) 2 . 

Part 3: In wells Bl though B8, add 5 drops of 0.1 mol/L Pb(N0 3 ) 2 . 

Part 4: In wells CI though C8, add 5 drops of 0.1 mol/L Ni{N0 3 ) 2 . 

Part 5: In wells Dl though D8, add 5 drops of 0.1 mol/L Co(N0 3 ) 3 . 

Part 6: In wells El though E8, add 5 drops of 0.1 mol/L Fe(N0 3 ) 3 . 

Part 7: In wells Fl though F8, add 5 drops of 0.1 mol/L AgN0 3 . 

Part 8: In wells Gl though G8, add 5 drops of 0.1 mol/L HN0 3 . 

Part 9: In wells Al through to Gl, add 5 drops of 0.1 mol/L Na 2 C0 3 . 

Part 10: In wells A2 though G2, add 5 drops of 0.1 mol/L Na 2 SOi. 

Part 11: In wells A3 though G3, add 5 drops of 0.1 mol/L NaCl. 

Part 12: In wells A4 though G4, add 5 drops of 0.1 mol/L Nal. 

Part 13: In wells A5 though G5, add 5 drops of 0.1 mol/L Na 2 CrO^. 

Part 14: In wells A6 though G6, add 5 drops of 0.1 mol/L Na 2 Cr 2 0^. 

Part 15: In wells A7 though G7, add 5 drops of 0.1 mol/L NaOH. 

Part 16: In wells A8 though G8, add 5 drops of 0.1 mol/L NH^OH. 

Part 17: Record all of your observations into your data table. 

Part 18: Clean Up. Empty the contents of the micro plate into the sink and rinse the plate and the sink 
with plenty of water. Wash your hands and the container thoroughly. 

Analysis 



1. Which reactions resulted in the formation of a precipitate? 

2. Write balanced chemical equations for the reactions found in question 1. Can you determine based 
on your observations in this lab what the precipitate is likely to be? If so, indicate that in your 
chemical reaction. 

3. What other physical changes were observed? 



367 



www.ckl2.org 



4. Write the balanced chemical equations for these reactions. 

5. What three questions do you have as a result of doing this experiment? 

Conclusion 

What conclusions did you make as a result of doing this experiment? 

Types of Chemical Reactions 

Investigation and Experimentation Objectives 




In this activity, the student will see the various types of chemical reactions and will associate particular 
observations with specific reaction types. 

Pre-lab Questions 

1. How does a decomposition reaction differ from the other types of chemical reactions? 

2. In a combination reaction, what other products form in addition to any new compound? 

3. Which two ions are bioaccumulative and should be used in very small quantities? 

4. During which steps (give numbers) of the procedure would you expect to produce a gas? 

5. What safety precaution applies when heating a test tube? 

Introduction 

In this experiment, you will learn to differentiate among five general types of chemical reactions. You 
will carry out certain representative reactions yourself, while others will be demonstrated by your teacher. 
From your observations you will attempt to identify the products of each reaction, and to determine the 
type of reaction that has taken place. The types of reactions that you will consider are the following: 
synthesis reactions, decomposition reactions, single replacement reactions, double replacement reactions, 
and combustion reactions. The majority of common chemical reactions can be classified as belonging to 
one of these categories. A brief description of each reaction type is provided below. 

(a) Synthesis reactions are reactions in which two or more substances combine to form a single product. 
The reactants may be elements or compounds, but the product is always a single compound. An example 
of a combination reaction is a reaction of sulfur trioxide and water to form sulfuric acid. 

SO m +H 2 (L) ^>H 2 S0 4 (aq) 

(b) Decomposition reactions are reactions in which a single substance breaks down into two or more 
simpler substances. There is always just a single reactant in a decomposition reaction. An example of a 
decomposition reaction is the breakdown of calcium carbonate upon heating. 

CaCO^fA + heat — > CaO^ + C0 2 (g) 

(c) Single replacement reactions are reactions in which an element within a compound is displaced to 
become a separate element. This type of reaction always has two reactants, one of which is always an 
element. An example of a single replacement reaction is the reaction of zinc metal with hydrochloric acid. 

Zit( s ) + 2 HClfa) -» ZnCI 2 ( aq ) + H 2 ( g ) 
www.ckl2.org 368 



(d) Double replacement reactions are reactions in which a positive ion from one ionic compound exchanges 
with the positive ion from another ionic compound. These reactions typically occur in aqueous solution 
and result in either the formation of a precipitate, the production of a gas, or the formation of a molecular 
compound such as water. An example of a double-replacement reaction is the reaction that occurs between 
aqueous silver nitrate and aqueous sodium chloride. A precipitate of solid silver chloride is formed in the 
reaction. 

AgN0 3(aq) + NaCl {aq) -> AgCl {s) + NaN0 2(aq) 

(e) Combustion reactions are reactions in which an element or compound reacts rapidly with oxygen gas 
to liberate heat and light energy. Commonly, the compounds combining with oxygen in these reactions are 
hydrocarbons, compounds containing hydrogen and carbon. The well-known combustible fuels kerosene 
and gasoline, for instance, are hydrocarbon mixtures. The complete combustion of a hydrocarbon yields 
carbon dioxide and water as well as reaction products. If insufficient oxygen is available, combustion will 
not be complete and carbon monoxide and elemental carbon may be obtained as additional products in 
the reaction. An example of a combustion reaction is the burning of methane gas to give water (in the 
form of steam), carbon dioxide, heat, and light. 

CH A{g) + 2 2(g) -» C0 2 ( g) + 2 H 2 (g) + heat + light 
Objectives: 

• To observe chemical reactions in order to determine the reaction type. 

• To write balanced chemical equations for each reaction. 

Apparatus and Materials 

Iron filings 

safety glasses 

Copper(II) sulfate pentahydrate 

2 x small test tubes 

Magnesium, turnings 

2 X medium test tubes 

0.1 mol/L copper(II) sulfate 

lx large test tube 

0.2 mol/L lead(II) nitrate 

1 x test tube holder 

0.2 mol/L potassium iodide 

gas burner 

3% hydrogen peroxide 

ring stand & clamp 

5 mol/L hydrochloric acid 

dropper pipette 

3% sulphuric acid (teacher demo) 

crucible tongs 

sodium bicarbonate (teacher demo) 

electrolysis apparatus (teacher demo) 

limewater (teacher demo) 

1-holed rubber stopper (teacher demo) 

369 www.ckl2.org 



• toothpicks 

• 1 glass tube, 25 cm long bent at 90° in 

• matches 

Procedure 

1. Draw a table similar to the one below to use for collecting your observations. 

Table 12.5: Observations 

Reaction Observations Reaction Type 

Fe and CuSO± 
Pb(N0 3 ) 2 and KI 
C11SO4 ■ 5H2O and heat 
Mg and HCl 
H2O2 and heat 
Electrolysis of H 2 
NaClOi and heat 



2. Iron metal and copper(II) sulfate solution. Fill a small test tube halfway with copper(II) sulfate 
solution. Add 2 g (about \ of a small test tube) of iron filings to the solution. Observe the reaction after 
5 minutes. Record your observations in the observation table. Discard the solid contents of the test tube 
into the waste container provided. The liquid portion can be poured down the sink. 

3. Lead(II) nitrate and potassium iodide solutions. Put 2 mL of lead(II) nitrate solution in the test 
tube. Add 5 to 10 drops of potassium iodide solution. Record your observations. Discard the contents of 
the test tube into the waste container and rinse the tube with water. 

4. Action of heat on copper(II) sulfate solution. Put two or three pea-sized crystals of copper(II) 
sulfate pentahydrate into a large, dry test tube. Fasten a utility clamp to the upper end of the test tube. 
Hold the tube by the clamp so that it is almost parallel with the surface of the lab bench. CA UTION: 
Do not point the open mouth of the tube at yourself or anyone else. Heat the crystals gently 
at the bottom of the tube (where the crystals are located) in a burner flame for approximately 30 seconds 
recording your observations. When the test tube has cooled, discard its contents into the waste container 
provided. 

5. Magnesium metal and hydrochloric acid. Fill one medium-size test tube halfway with 6 mol/L 
hydrochloric. CAUTION: Hydrochloric acid is corrosive. Place the test tube in the test tube rack. 
Put several pieces of magnesium turnings into the acid solution. If you observe a gas forming, test for 
its identity by holding a burning wood splint at the mouth of the test tube. Do not put the splint into 
the solution. Record your observations. Decant the liquid portion of the test tube contents into the sink; 
discard the solid into the waste container provided. 

6. Action of heat on hydrogen peroxide. Add 2 mL of the 3% hydrogen peroxide solution to a medium 
test tube. Use a utility clamp to secure the tube to a ring stand. CAUTION: Make sure that the 
mouth of the tube is pointed away from you and away from everyone else. Heat the solution 
very gently. If you observe a gas forming, test for its identity by inserting a glowing wood splint at the 
mouth of the test tube. Do not put the splint into the solution. Record your observations. Rinse the 
contents of the test tube into the sink. 

TEACHER DEMONSTRATIONS 

7. Action of electricity on water (Electrolysis). Water can be broken down to its component 

www.ckl2.org 370 



elements by passing electricity through it. This process is called electrolysis. The apparatus used for 
this demonstration will be explained by your teacher. Make your observations of the reaction at several 
intervals during a period of 5 - 10 minutes. 

8. Action of heat on sodium bicarbonate. Solid sodium bicarbonate will be heated strongly in a 
test tube for 2 minutes. The gas that is given off will be tested by exposing it to a burning splint, and 
by bubbling it through limewater (a saturated solution of calcium hydroxide, Ca{OH)2). Record your 
observations of these tests. 

Data Analysis 

1. Decide which type of reaction is represented by each reaction observed in this experiment. Record 
your answers in your observation table. 

2. Write a balanced chemical equation for each chemical reaction observed. 

3. Although you did not work with any synthesis reactions in this experiment, can you describe one or 
give an example of one that you might have seen before or read about. Write a balanced equation 
for this reaction. 

Results and Conclusions 

1. Describe in your own words the five types of chemical reactions that were discussed in the introduction 
to this experiment. Explain how each type of reaction can be identified. 

2. List the tests that were used in this experiment to identify gases. 

Extension 

1. Make a list of the reactions observed in previous experiments. Identify the types of reaction in as 
many cases as possible. 



Image Sources 



371 www.ckl2.org 



Chapter 13 

The Liquid State 



13.1 The Properties of Liquids 

Lesson Objectives 

• The student will explain the basic behavior and characteristics of liquids using the molecular arrange- 
ment in liquids. 

Introduction 

Liquids and solids differ from gases in two important ways. The particles (atoms, molecules, or ions) are 
much closer together so the total volume of a liquid or solid is much closer to the sum of the volumes of 
the molecules. Also, attractive forces between the particles in liquids and solids are much stronger. The 
strength of these forces of attraction between particles is a major reason that the substance is in the liquid 
or solid state rather than gaseous. 

Liquids Maintain Their Volume But Take the Shape of Their 
Container 

The arrangement of the molecules in a liquid structure accounts for most of the physical properties of 
liquids (see Figure 13.1). 



5 



S AS 



& W ysft 



LIQUID 



SOLID 



Figure 13.1: The Three Phases of Matter. 

Liquids have much less space between molecules and stronger attractive forces than gases. In the liquid 
state, attractive forces between molecules are a major factor in the behavior of the liquids. Although liquid 
structure has small spaces so the molecules can move past one another, the attraction between particles 



www.ckl2.org 



372 



keep them from getting very far apart. Therefore, liquids maintain their own volume but take the shape 
of their container. 100 mL of liquid will be 100 mL in any container but because the liquid molecules are 
not held in a tightly packed pattern (like solids), the molecules can move past one another and so liquids 
will flow to fit the shape of their container 



Liquids Have Greater Densities Than Gases 

In gases, the distance between the molecules is so great that the size of the molecules themselves become 
inconsequential. A gas is considered to be essentially empty space. (If the picture above were realistic 
of gas structure, the molecules would be so small you couldn't see them.) If we consider the volumes of 
6.02 X 10 23 molecules of oxygen gas (O2) and 6.02 X 10 23 molecules of Freon gas (CF4) both at STP, we 
know that the volumes will be the same, 22.4 liters. This is because the sizes of the molecules themselves 
are negligible compared to the empty space in a gas. The fact that the Freon molecules are several times 
larger than the oxygen molecules makes no difference. In a gas, you are measuring the volume of the empty 
space. If we consider the volume of 6.02 x 10 23 molecules of liquid oxygen and 6.02 x 10 23 molecules of 
liquid Freon both at STP, we find the volume of the oxygen is about 28 mL and the volume of the Freon 
is about 55 mL. In the case of liquids, the volumes of the molecules themselves make a difference. In 
liquids, the volume of a group of molecules is related to the volume of the individual molecules and so 
equal moles of liquids DO NOT occupy equal volumes under the same conditions. Since liquids have many 
more molecules in a smaller volume, they will have much greater densities than gases. 



Liquids Are Almost Incompressible 

When a substance is compressed, it is not the molecules themselves that are compressed; it is the space 
between the molecules that is compressed. Gases have lots of empty space and so are easily compressed. 
A pressure of 3.0 atm compresses a gas to one-third its volume at 1.0 atm. Liquids have very little space 
between molecules and do not compress easily - a pressure of 3.0 atm will have virtually no effect on 
the volume of a liquid. Liquids are used in hydraulic systems because of their ability to flow to fit their 
container and their incompressibility. 



Liquids Diffuse More Slowly Than Gases 

Diffusion in gases (mixing) is nearly instantaneous. If you release a colored gas in a container of non-colored 
gas, the color spreads evenly throughout the container in a second or two. In liquids, diffusion is a much 
slower process. The smaller spaces to move through and the almost constant collisions cause the molecules 
to require much more time to move from one side of the container to the other. 



Lesson Summary 

• Molecules in the liquid phase have some freedom of movement but their motion is much more re- 
stricted than that of gases. 

• Liquids are only slightly compressible. 

• Liquids diffuse more slowly than gases. 

373 www.ckl2.org 



Review Questions 

1. The molar volumes of solid silicon and solid bromine under the same conditions are different and 
the molar volumes of liquid silicon and liquid bromine under the same conditions are different, but 
the molar volumes of gaseous silicon and gaseous bromine under the same conditions are exactly the 
same. Explain. 

Vocabulary 

incompressible The terms compressibility and incompressibility describe the ability of molecules in a 
fluid to be compacted (made more dense). 

13.2 Forces of Attraction 

Lesson Objectives 

• The student will identify liquids whose intermolecular forces of attraction are due to London disper- 
sion forces, polar attractions, and hydrogen bonding. 

• The student will describe some of the unique properties of water that are due to hydrogen bonding. 

• The student will select from comparative compounds, the ones most likely to form hydrogen bonds. 

• The student will select from comparative compounds whose intermolecular forces are London disper- 
sion forces, the one most likely to have the strongest intermolecular forces. 

Introduction 

Forces on the molecular level are divided into two categories - the forces inside a molecule holding atoms 
together to form the molecule and the forces between molecules holding the molecules in the liquid or solid 
state. The forces inside the molecule are called intramolecular forces and the forces between molecules are 
called intermolecular forces. You can relate them to bus systems where intracity buses move people around 
inside one city and intercity buses move people from one city to another. 

Intramolecular Forces 

Intramolecular forces refer to the forces inside a molecule that holds the atoms together to form molecules. 
Essentially, there is only one force in this category and that is covalent chemical bonds. See below figure. 

The only intramolecular forces are covalent chemical bonds. (Source: Richard Parsons. CC-BY-SA) 



\ • I 



Even though we sometimes write ionic compounds in equations and call the equations "molecular equa- 
tions", ionic compounds are not considered to form molecules. Ionic compounds in the solid state from 
crystal lattices in which each ion is equally attracted to more than one oppositely charged ion. The only 
true molecules are in covalently bonded compounds. 

www.ckl2.org 374 



Intermolecular Forces of Attraction 

We will consider three types of intermolecular forces of attraction. These are the forces that hold molecules 
in the condensed states (liquid and solid). See below figure. 

Intermolecular forces of attraction are the forces between molecules that hold them in the liquid or solid 
state. (Source: Richard Parsons. CC-BY-SA) 



H *v _ 



H 



When a substance changes phase as in a solid melting to liquid or a liquid vaporizing to gas, the inter- 
molecular forces must be overcome. Therefore, the stronger the intermolecular forces of attraction, the 
greater the molecular motion (temperature) that will be required to overcome them. The liquids with the 
strongest intermolecular forces of attraction will have the highest boiling points and vice versa. 

London Dispersion Forces 

The weakest type of intermolecular force is called London dispersion forces. London dispersion forces 
occur between all atoms and molecules but they are so weak, they are only considered when there is no 
other intermolecular force. For example, London dispersion forces exist between water molecules but water 
molecules also have a permanent polar attraction so much stronger than the London dispersion forces that 
the London dispersion force is insignificant and not mentioned. 

The cause of London dispersion force is not obvious. Although we usually assume the electrons of an atom 
are uniformly distributed around the nucleus, this is apparently not true at every instant. As the electrons 
move around the nucleus, at a given instant, more electrons may be on one side of the nucleus than the other 
side. This momentary nonsymmetrical electron distribution can produce a temporary dipolar arrangement 
of charge. This temporary dipole can induce a similar dipole in a neighboring atom and produce a weak, 
short-lived attraction. 

The cases where London dispersion forces would be considered as the intermolecular force of attraction 
would be in the noble gases and non-polar molecules such as H2,N2,Cl2,CH4,CCU,CC>2,S F§ and so forth. 
Since non-polar molecules do not have a permanent dipole and no further bonding capacity, their only means 
of attracting each other is through London dispersion forces. Some of the substances whose intermolecular 
forces of attraction are London dispersion forces are held in the liquid state very weakly and therefore, 
would have the lowest melting points of all substances. (See Table 13.1). 

Table 13.1: Boiling Points of Some London Dispersion Forces Liquids 

Substance Boiling Point (°C) 

Helium, He -269.7 

Neon, Ne -248.6 

Argon, Ar -189.4 

375 www.ckl2.org 



Table 13.1: (continued) 



Substance Boiling Point (°C) 

Krypton, Kr -157.3 

Xenon, Xe -111.9 

Hydrogen, H2 -253 

Oxygen, 2 -182 

Methane, CH A -161 

Carbon Dioxide, CO2 -78 



The temporary dipoles that cause London dispersion forces are affected by the molar mass of the particle. 
The greater the molar mass of the particle, the greater is the force of attraction caused by London dispersion 
forces. The molar masses of H2, N2, and O2 are 2, 28, and 32 g/mol, respectively, and their boiling points 
increase in similar fashion; H2 is -253°C, N2 is — 196°C, and O2 is -183°C. When the molar masses of 
London dispersion force liquids become high enough, the substances will be liquid even at room temperature 
and above. Carbon tetrachloride, molar mass 154 g/mol, and bromine, molar mass 160 g/mol, boil at 
+77°C and +59°C, respectively. Many long carbon chain, non-polar substances such as gasoline and oil 
remain liquids at common temperatures. 

Polar Attractions 

When covalent bonds form between identical atoms such as in H2, N2, O2, and so on, the electrons shared 
in the bonds are shared exactly equally. The two atoms have the same electronegativity and therefore, the 
same pull on the shared electrons. (See Figure 13.2). 





Center of Negative Charge Center of Positive Charge 



Figure 13.2: The centers of positive and negative charge. 

The center of negative charge for the entire molecule will be in the exact center of the molecule. This will 
coincide with the center of positive charge for the molecule. When the center of negative charge and the 
center of positive charge coincide, there is no charge separation and no dipole. 

If the two atoms sharing the bonding pair of electrons are not of the same element, the atom with the greater 
electronegativity will pull the shared electrons closer to it. Because of the resulting uneven distribution of 
electrons, the center of negative charge will not coincide with the center of positive charge and a dipole 
is created on the molecule. When the centers of positive and negative charge do not coincide, a charge 
separation exists and a dipole is present. 

The end of the molecule with the more electronegative atom will have a partial negative charge and the 
end of the molecule with the more electropositive atom will have a slight positive charge. The symbols 8+ 
and 6- are used because these are not full 1+ and 1- charges. See below figure. 

The polar molecule. (Source: Richard Parsons. CC-BY-SA) 



www.ckl2.org 376 




This polarity is much less of a charge separation than in an ionic bond. These charges are only fractions 
of full 1+ and 1- charges. How much polarity a bond will experience depends on the difference in the 
electronegativities of the atoms. 

In the case of a symmetrical molecule with polar bonds, the symmetry of the electron displacements of 
two or more electron pairs will keep the center of negative charge in the center of the molecule coincident 
with the center of positive charge and so no dipole will occur. See below figure. 

When the electron shift in polar bonds is symmetrical over the entire molecule, there will be no dipole. 
(Source: Richard Parsons. CC-BY-SA) 



:0 = C = 

■ ■ 

For example, in the CO2 molecule, both carbon-oxygen bonds are polar, but the shift of bonding electrons 
toward the oxygen is the same on both sides of the carbon atom and the center of negative charge remains 
in the center. 

For molecules that have a permanent dipole, the attraction between oppositely charged ends of adjacent 
molecules are the dominant intermolecular force of attraction. 

Figure 13.3: Polar attractions in a solid hold the molecules together. 

Figure (13.3) represents a polar solid, a polar liquid would look similar except the molecules would be 
less organized. On the average, these polar attractions are stronger than London dispersion forces so 
polar molecules, in general, have higher boiling points than London dispersion liquids. There is significant 
overlap, however, between the boiling points of the stronger London dispersion molecules and the weaker 
polar molecules. 

The organization of a substance composed of polar molecules depends on the competition between the 
strength of the polar attractions and the molecular motion of the molecules. At higher temperatures, the 
molecular motion of the molecules is strong enough to disrupt the polar attractions but at low tempera- 
tures, the molecular motion is reduced and the polar attractions can hold the molecules in a structured 
arrangement. In liquid and gaseous forms, the molecules can also turn freely. If a pair of polar molecules 
in liquid form are aligned positive end to positive end, it is no problem for one of the molecules to spin 
around and align its negative end to the positive end of the other molecule. (See Figure 13.4). 

This turning of polar molecules toward an attractive force can be seen in a macroscopic situation as well. 

377 www.ckl2.org 



non-polar liquid 



water (polar liquid) 



charged red 



/ 




charged red 



Figure 13.4: A stream of polar liquid will bend toward a charged rod. 

If we allow a very thin stream of water to run from a faucet and we bring a charged object (rubber comb 
run through hair, balloon rubbed on wool sweater, etc) near the stream of liquid, the stream will bend 
its path toward the charged object. It doesn't make any difference if the charged object is positively or 
negatively charged because the water molecules in the stream will turn their oppositely charged end toward 
the charged object. In the sketch at right, the path of a non-polar liquid is not deflected by the charged 
rod but the path of the water stream is deflected by the charged rod. 

Hydrogen Bonds 

There are several polar molecules whose polar bonds are so strong they merit separate attention. These are 
the polar attractions that occur in molecules where hydrogen is bonded to nitrogen, oxygen, or fluorine. 
The polar attractions in these molecules are nearly 10 times as strong as regular polar attractions. These 
extra strong polar attractions that occur with the H — N, H — O, and H — F bonds are called hydrogen 
bonds which distinguishes them from regular polar attractions but keep in mind that they are, in fact, 
polar attractions, albeit very strong ones. 

There is more than one explanation for why these three combinations form hydrogen bonds. There are 
only ten elements that have greater electronegativity than hydrogen and only four that have a significantly 
greater electronegativity than hydrogen. Three of the elements that have significantly greater electroneg- 
ativity than hydrogen are nitrogen, oxygen, and fluorine - the three elements that form hydrogen bonds 
in compounds with hydrogen. When hydrogen bonds with atoms whose electronegativities are less than 
or equal to the electronegativity of hydrogen, the other atom cannot pull the shared electrons away from 
hydrogen. (See Figure 13.5). 

Therefore, the hydrogen nucleus (a single proton) is shielded by electrons in its electron cloud. When 
hydrogen chemically bonds with nitrogen, oxygen, or fluorine, the very high electronegativities of these 
atoms CAN pull the electrons far away from the hydrogen atom, thus removing the shielding electrons 
from the proton nucleus of hydrogen. When the polar attractions between HF molecules or other hydrogen 
bonding compounds set up, the negative end of a molecule can get very close to the proton on the positive 
end of another molecule because there are no electrons for shielding. The closeness of the charges causes 
the extra strong polar attractions in these compounds. The characteristics of a liquid that forms hydrogen 
bonds are significantly different from similar compounds that do not form hydrogen bonds. In the case of 
water, this is very important to life on earth. 



www.ckl2.org 



378 



In HBr. the bonding electrons remain 
partially around the hydrogen and 
shield its nucleus 



© 



o 



In HF. the bonding electrons are 
pulled away leaving an unshielded 
nucleus. 



Figure 13.5: With highly electronegative non-metals, the electron shift away from the hydrogen atom leaves 
an unshielded proton. 

A homologous series of compounds are compounds of the elements of a family, each bonded to the same 
other element. For example, family 4A in the periodic table consists of carbon, silicon, germanium, and 
tin. If each of these is bonded to hydrogen, it would produce a homologous series, C//4, SiH^, GeH^, and 
SnH^. If we graph the boiling points of this homologous series, we would get the graph sketched in Figure 
13.6. The large majority of graphs of the boiling points of homologous series would look like this in that 
the boiling points increase as molar mass increases. If we graph the boiling points of the homologous series 
of family 6A combined with hydrogen, we get a quite different graph. 




■200 



100 



B.P. in C 



60 120 

Molar Mass 



100 




60 120 

Meier Mass 



Figure 13.6: Boiling point comparisons: family 4A hydrogen compounds to family 6A hydrogen compounds. 

The higher molar mass compounds in the series follow the normal pattern of decreasing boiling points as 
the molar mass decreases. But, when we get to water, suddenly the boiling point is way out of line - it 
is more than 150 degrees too high. The explanation for why liquid water is held together far more tightly 
than expected - hydrogen bonds. Graphs of the boiling points of the homologous series of hydrogen with 
5A family members and hydrogen with 7A family members would be similar. The boiling point of NH3 
and HF are greatly different from what would be expected. (See Figure 13.7). 

The fact that water forms hydrogen bonds has effects so large that it is impossible in this text to delineate 
them all. We will consider a few but there are many more not in this list. 

1. The normal boiling point of water is 100°C If water did not form hydrogen bonds and instead had 
a regular polar attraction between molecules, its boiling point would be somewhere around — 60°C. 
The average temperature of the surface of the earth is 15°C If water did not form hydrogen bonds, 
the oceans and lakes would vaporize and earth would not be a water planet. Therefore, it would 
likely not be a planet with life on it. 

2. If you gathered 100 substances of all types (ionic, metals, regular polar substances, London dispersion 



379 



www.cki2.0rg 





Water with no significant hydrogen bonding. Water with significant hydrogen bonding. 

Figure 13.7: Significant hydrogen bonding causes water molecules to line up end-to-end. 



substances) and water and you arranged to have each of these substances in both solid and liquid form 
and you dropped pieces of solid into the corresponding liquid, you would find exactly one compound 
whose solid floated in the liquid. For all the other compounds, the pieces of solid would sink to 
the bottom of the liquid. Only water would have the solid floating on the liquid. The reason for 
this is that as you cool the liquid of almost all substances to freeze them, the substance contracts 
as it cools - the molecules move around slower and intermolecular forces pull the molecules closer 
together. As a result, the solids are denser than the liquids and the solids will sink in the liquids. 
Water, of course, is the exception. When water cools to its freezing point and solidifies, it expands. 
When water molecules are at or above 4°C, the molecular motion is sufficient to keep the hydrogen 
bonds from locking water molecules into an end-to-end molecular complex with large holes in the 
structure. When water is cooled below 4°C, the molecular motion is inadequate to break up this 
complex structure - so water molecules begin forming end-to-end chains and the water expands 
because of the holes in the structure. Therefore, solid water is less dense than liquid water and ice 
floats on water. You have seen ice cubes floating in water a thousand times and you have probably 
never seen any other substance where you see the solid interacting with the liquid, so the ice floating 
looks normal to you. If you were familiar with the behavior of all other substances in this situation 
and then you saw someone place ice cubes in water for the first time, you would probably think they 
had just done a magic trick. One of the effects of the fact that ice floats on water is that when cold 
weather comes to areas in the northern and southern parts of the earth, the cold air contacts the 
water on the surface of lakes and freezes that surface water. If water were like other substances and 
the solid sank to the bottom, then the cold air would freeze the new surface and it would sink and 
so on until the entire lake would be frozen top to bottom. No water dwelling animals would survive 
such an occurrence. But, actually, when the cold air freezes the surface of a lake, the ice floats, stays 
on top and insulates the rest of the rest of the water from the cold air. Only the surface freezes and 
the animals that live in the water survive the winter. 

3. One of the factors in the weathering of rocks in geographical areas that have winter is that rain water 
enters rocks through cracks and then freezes and expands, fracturing the rock. 

4. Some biologically active molecules such as DNA require a particular shape for their function. If you 
think of a long-chain molecule as something similar to a string, how can such a molecule have and 
hold a shape. At points along its length, the molecule can be linked to itself with different types of 
attractions - one of which are hydrogen bonds. See below figure. 

5. Hydrogen bonds form links to maintain shapes of molecules. (Source: Richard Parsons. CC-BY-SA) 

www.ckl2.org 380 



6. 



7. As long as water is above 4°C, when it is cooled, like other substances, it contracts and becomes 
denser. When water passes through 4°C, significant hydrogen bonding begins to form and the water 
expands and becomes less dense. The maximum density for water is at 4°C. Many animals that 
live in water and require oxygen, use oxygen that is dissolved in the water. Water in lakes becomes 
oxygenated (has dissolved oxygen) by the action of wind and waves at the surface. For deep lakes, 
diffusion is inadequate to move the oxygenated water to the bottom of the lake. Once the oxygen 
was used up in the water at the bottom of a lake, oxygen using animals would be unable to live there. 
For lakes in northern climates, the surface passes through the temperature 4°C twice per year, once 
as it cools in the fall and once as it warms in the spring. During these two times, the water at the 
surface would sink to the bottom because it is denser and the sinking water would be oxygenated. 
These periods are called spring and fall "turnovers." The turnovers provide oxygenated water at the 
bottom of the lakes. 

The special chemical properties of water are explored, along with the need for its protection and conserva- 
tion. Water (http : //www . learner . org/vod/vod_window . html?pid=804) 

Ionic Liquids 

Ionic compounds may also exist in liquid form. The intermolecular forces of attraction in ionic liquids 
would, of course, be the electrostatic attraction between oppositely charged ions. The charges on ions are 
complete charges of 1+, 2+, 1-, 2- and so on. The attractions due to these charges are much greater than 
those of polar molecules, even the especially strong polar attractions of hydrogen bonds, and so the boiling 
points of ionic liquids would be much higher than the molecular liquids discussed previously. Sodium 
chloride, for example, boils at 1465° C. 

Metallic Liquids 

Metals may also exist in the liquid state. The bonding and therefore, the intermolecular forces of attraction 
for metals will be covered in the next chapter. 

Lesson Summary 

• Molecules are held together in the liquid phase by intermolecular forces of attractions. 

• London dispersion forces are a very weak intermolecular force of attraction caused by a temporary 
electrostatic attraction between the electrons of one molecule or atom and the nucleus of another. 

• Polar attractions are an intermolecular force of attraction caused by the electrostatic attraction 
between permanent dipoles that exists on polar molecules. 

381 www.ckl2.org 



• Hydrogen bonds are an exceptionally strong type of polar attraction that occurs between molecules 
that have H — F bonds, H — O bonds, or H — N bonds. 

• Ionic substances also exist in liquid form; the attractions between the ions are the attractions between 
ionic charges and are much stronger than polar attractions or hydrogen bonds. 

Review Questions 

1. Which of the following molecules would you expect to have the higher melting point? Why? 

(a) CH A or H 2 S 

(b) H 2 or H 2 S 

(c) HCl or Cl 2 

(d) Nal or NH 3 

(e) S F4 or C//4 

2. The structures of vitamins C and E are shown above. Which of the following statements is correct? 





(a) Vitamin E has more opportunities for hydrogen bonds than vitamin C. 

(b) The melting point of vitamin E is likely to be higher than that of vitamin C. 

(c) Vitamin C is likely to be very soluble in a non-polar solvent. 

(d) Vitamin C should have a higher solubility in water than vitamin E. 

(e) Vitamin C would be described as a "fat-soluble" vitamin. 

Vocabulary 

hydrogen bond The exceptionally strong polar attraction between a hydrogen atom in one molecule 
and a highly electronegative atom (N, O, F) in another molecule. 

London dispersion forces Electrostatic attractions of molecule or atoms for nearby atoms or molecules 
caused by the temporary unsymmetrical distribution of electrons in electron clouds. 



13.3 Vapor Pressure 

Lesson Objectives 



• The student will be able to describe the processes of evaporation and condensation. 

• The student will be able to describe vapor pressure equilibrium. 

Introduction 

The phase of a substance is essentially the result of two forces acting on the molecules. The molecules of 
a substance are pulled together by intermolecular forces of attraction that were discussed in the previous 

www.ckl2.org 382 



section. Some of these intermolecular forces are weak and some are strong. The molecules of a substance 
also have kinetic energy so that the molecules are in constant random motion and are in almost constant 
collisions with each other. The motion and collisions of molecules push them away from each other. 
Without intermolecular forces of attraction, the molecules of all substances would move away from each 
other and there would be no condensed phases (liquids and solids). 

If the forces caused by molecular motion are much greater than and dominate the intermolecular forces of 
attraction, the molecules will separate and the substance will be in the gaseous state. If the intermolecular 
forces of attraction are dominantly stronger than the molecular motion, the molecules will be pulled into 
a closely packed pattern and the substance will be in the solid state. If there is some balance between 
molecular motion and intermolecular forces of attraction, the substance will be in the liquid state. 

When substances are heated or cooled, their average kinetic energy increases or decreases, their molecular 
motion increases or decreases, and the substance may change phase. A substance in the solid phase 
(intermolecular forces dominate) can be heated until the molecular motion balances the intermolecular 
forces and the solid will melt to liquid. The liquid may be heated until the molecular motion completely 
overcomes the intermolecular forces and the liquid will vaporize to the gaseous state. 

When you open a jar of perfume, your nose detects the substance almost immediately. You can see that 
the substance is in liquid form and it is not boiling, yet some of that material obviously entered the gaseous 
state and reached your nose. 

Evaporation 

The temperature of a beaker of water is a measure of the average kinetic energy of the molecules in the 
beaker. That does not mean that all the molecules in the beaker have exactly the same kinetic energy. 
Most of the molecules will be within a few degrees of the average but a few molecules may be considerably 
hotter or colder than the average. The kinetic energy of the molecules in the breaker will have a distribution 
curve similar to a standard distribution curve for most naturally occurring phenomena. (See Figure 13.8). 



Number of 
Occurrences 




Vorioble 



A standard distribution curve like this could represent the length of 
grasshopper legs in a random sample of grasshoppers or the 
distribution of 1Q for a random sample of humans. 

Figure 13.8: Standard Distribution Curve 

Naturally occurring phenomena usually have most of the instances near the average and the number of 
instances become less as the value gets farther from the average. 

In the case of a beaker of water, some of the molecules will have an average temperature below the boiling 
point and some of the molecules will have a temperature above the boiling point (see Figure 13.8). The 
dashed yellow line is the average temperature of the molecules and would be the temperature shown on a 

383 www.ckl2.org 





y*X^ 




Number of 


/\ 




Molecules 


/ \ 








B.P. 

-, ^> ■ ■ I — 



Temperature 



Figure 13.9: The distribution of temperature among the molecules of a beaker of water. 



thermometer inserted into the liquid. The red line represents the boiling point of water (100°C at 1.00 atm 
pressure) and the area under the curve to the right of the red line represents the molecules of water that 
are above the boiling point. In order for a molecule of liquid that is above the boiling temperature to 
escape from the liquid, it must either be on the surface or it must be adjacent to many other molecules 
that are above the boiling point so that the molecules can form a bubble and rise to the surface. See below 
figure. 

Water boils only when a sufficient number of adjacent molecules are above the boiling point and can form 
bubbles of gaseous water. (Source: Richard Parsons. CC-BY-SA) 




The only circumstance when there are enough molecules above the boiling point to form bubbles is when 
the average temperature is at the boiling point. For single molecules above the boiling point, they must 
wait until their random motion gets them to the surface and then the molecule can leave the liquid and 
enter into the gaseous phase. This process of molecules escaping to the gaseous phase from the surface of 
a liquid when the average temperature of the liquid is below the boiling point is called evaporation. 

The process of phase change is a little more complicated than just having the molecules reach the boiling 
point. Objects that attract each other and are separated have potential energy due to that attraction 
and separation. An object held above the earth is attracted to the earth by earth's gravity and has 
potential energy due to the attraction and separation. The amount of potential energy can be calculated 
by multiplying the force of attraction times the distance of separation. Two oppositely charged objects 
that are separated have potential energy. Objects with opposite magnetic poles that are separated have 
potential energy. A stretched rubber band has potential energy. Gaseous molecules that have a force of 
attraction between them but are separated have potential energy. Molecules in the liquid state and the 
same molecules in the gaseous state at the same temperature DO NOT have the same total energy. If they 
are at the same temperature, they have the same kinetic energy but the gaseous molecules have potential 
energy that the liquid molecules do not. Molecules in the liquid state that are hot enough to exist in the 
gaseous state must absorb energy from their surroundings to provide that potential energy as they change 
phase. This potential energy is called the heat of vaporization and it is absorbed by evaporating or 



www.ckl2.org 



384 



vaporizing molecules from the kinetic energy of the liquid. 

You are at least somewhat familiar with evaporation. You know that if you leave a saucer of water sitting 
out on the countertop, the water will slowly disappear - and yet, at no time is the temperature of the 
water ever at the boiling point. The water in an open container continues to evaporate until it is all in the 
vapor state. When molecules of a liquid are evaporating, it is clear that it is the hottest molecules that are 
evaporating. It might seem that once the molecules whose temperature was above the boiling point are 
gone, no more evaporation would occur. Here's the reason evaporation continues. The temperature of the 
liquid is the average temperature of all the molecules. When the hottest molecules evaporate, the average 
temperature of those molecules left behind is lower and the molecules left behind have also contributed to 
the heat of vaporization to the evaporating molecules. The process of evaporation causes the remaining 
liquid to cool significantly. Heat flows from warmer objects to colder objects and so when the liquid cools 
due to evaporation, the surroundings will give heat to the liquid thus raising its temperature back up equal 
to the surroundings thus producing more hot molecules. This process can continue in an open container 
until the liquid is all evaporated. 

Many years ago, when people lived without electricity, they figured out that if they placed the butter dish 
on the dinner table in a shallow container of water, evaporation would cool the water and therefore the 
butter dish, enough to keep the butter from melting to a liquid. They also knew that if you put a container 
of water or milk in a fabric sack, soaked the sack in water, and swung the sack around in the air, the 
evaporation of the water from the sack would cool it and cool the container of milk or water to make it 
nicer to drink. Many hikers today use fabric canteen holders that they soak in water while hiking so that 
the water in the canteen will be cooler when they drink it. 

The rate of evaporation is related to the strength of the intermolecular forces of attraction, to the surface 
area of the liquid, and to the temperature of the liquid. As the temperature of liquids get closer to the 
boiling point, more of the molecules have temperatures above the boiling point and so evaporation is faster. 
Substances with weak intermolecular forces of attraction evaporate more quickly than those with strong 
intermolecular forces of attraction. Substances that evaporate readily are called volatile and those that 
hardly evaporate at all are called non- volatile. 

Condensation 

Liquids will usually evaporate to dryness in an open container. What happens, however, if the container is 
closed? When a lid is put on the container, the molecules that have evaporated are now kept in the space 
above the liquid. This makes it possible for a gaseous molecule to collide with another molecule or a wall 
and condense (the gas to liquid phase change) back to liquid. Molecules at the boiling point can exist in 
either liquid phase or gaseous phase - the only difference between them is the amount of potential energy 
they hold. See below figure. 

Evaporation and condensation both occur in a closed container. (Source: Richard Parsons. CC-BY-SA) 




For a liquid molecule with adequate temperature to exist in the gaseous phase, it is necessary for the 

385 www.ckl2.org 



molecule to gain the heat of vaporization. It does this by collision with adjacent molecules. For a gaseous 
molecule to return to the liquid phase, it must give up the same amount of potential energy that it gained. 
That amount of potential energy is called the heat of vaporization when it is being gained and it is 
called the heat of condensation when it is being lost - but it is exactly the same amount of energy. 

As more and more molecules evaporate in a closed container, the partial pressure of the gas in the space 
above the liquid increases. The rate at which molecules evaporate is determined by the temperature, the 
surface area, and what substance is involved. Once the substance, the surface area, and the temperature 
are established, the rate of evaporation will be constant. The rate at which the gas condenses is determined 
by the partial pressure of the gas, the surface area, and what substance is involved. Once the substance 
and surface area is established, the rate of condensation will only vary depending on the partial pressure of 
the gas. As the partial pressure of the gas in the space above the liquid increases, the rate of condensation 
will increase. 

In the section on evaporation, it was pointed out that as a liquid evaporates, the remaining liquid cools 
because the hottest molecules are leaving so the average decreases and the heat of vaporization is being 
absorbed from the remaining molecules. For similar reasons, when a gas is undergoing condensation, the 
temperature of the remaining gas increases because the coolest molecules are condensing, thus raising the 
average of those left behind, and the condensing molecules must give up the heat of condensation. 

Vapor Pressure Equilibrium 

You can follow the progress of the two activities (evaporation and condensation) in a thought experiment. 
Suppose we place some liquid water in an Erlenmeyer flask and seal it. No water has evaporated yet so 
the partial pressure of water vapor in the space above the liquid is zero. As a result, there will be no 
condensation. As the water evaporates (at a constant rate since the temperature and surface area are 
constant), the partial pressure of the water vapor increases. Now that some vapor exists, condensation 
begins. Since the partial pressure of the water vapor is low, the rate of condensation will be low. Over 
time, more and more water evaporates and the partial pressure of the water vapor increases. Since the 
partial pressure increases, the rate of condensation increases. Eventually, the rate of condensation will 
become high enough that it is equal to the rate of evaporation. Once this happens, the rate of molecules 
of water going into the vapor phase and the number of molecules condensing back to liquid are exactly 
the same and so the partial pressure no longer increases. When the partial pressure of the water vapor 
becomes constant, the rate of condensation is constant and is exactly EQUAL to the rate of evaporation. 
A condition called vapor pressure equilibrium has been established. As time goes on from this point, 
the amount of liquid cannot change, the amount of gas cannot change; neither the rate of evaporation 
nor the rate of condensation can change. Everything remains exactly the same, BUT the two activities 
continue. Evaporation continues and condensation continues at exactly the same rate. Each different liquid 
at each temperature will have an exact partial pressure of vapor that will be present when vapor pressure 
equilibrium is established. The pressure of the vapor in the space above the liquid is called the vapor 
pressure of that liquid at that temperature. (See Table 13.2). 

Table 13.2: Vapor Pressure of Water at Various Temperatures 

Temperature in °C Vapor Pressure in Torr 

4.6 

10 9.2 

20 17.5 

30 31.8 

40 55.3 

50 92.5 

www.ckl2.org 386 



Table 13.2: (continued) 



Temperature in °C 



Vapor Pressure in Torr 



60 
70 
80 
90 
100 



149.4 
233.7 
355.1 
525.8 
760.0 



Volatile liquids would have higher vapor pressures than water at the same temperature and non-volatile 
liquids would have lower vapor pressures at the same temperature. The amount of volume for the space 
above the liquid makes no difference. The partial pressures of the gases will reach the equilibrium value 
- if the space is small, it will take little gas to produce the pressure and if the space is large, it will take 
much more gas to produce the pressure. As long as you introduce enough liquid into the container so that 
vapor pressure equilibrium will be reached, then the precise vapor pressure will be attained. 

You might have noticed a subtle switch in vocabulary sometimes referring to the substance in the gaseous 
state as a gas and sometimes as a vapor. Chemists have agreed that a substance in the gaseous phase 
at temperatures above the boiling point of its liquid should be called a gas and if the temperature of the 
substance is below the boiling point, it should be called a vapor. 

You should also note that the equilibrium vapor pressure of a liquid is the same regardless of whether or 
not another gas is present in the space above the liquid. If the space above liquid water contains air at 
760 torr, and the liquid water evaporates until its equilibrium vapor pressure (25 torr) is reached, then the 
total pressure in the space above the liquid will be 785 torr. The presence of the air in no way affects the 
vapor pressure. 

Vapor Pressure Correction 

When gaseous substances are produced from chemical reactions and collected in the laboratory, they are 
usually collected over water. The "collection over water" technique is inexpensive and allows gaseous 
substances to be collected without having air mixed in. The process involves filling a collecting jar with 
water and inverting the jar in a pan of water without letting any water out or air in. 




Figure 13.10: Gas Collection Over Water. 

In Figure 13.10, the picture on the far left represents the collecting jar full of water and inverted in a pan 
of water. A tube runs from the reaction vessel where the gas is produced and is tucked under the edge of 
the collecting jar. As the gas is produced and comes out the end of the tube, it bubbles up through the 
water and pushes the water out of the jar. When the water in the collecting jar and the pan are exactly 
level, as in the picture at the far right, the pressure inside the collecting jar and the atmospheric pressure 
in the lab are equal. Using the pressure in the lab and the temperature in the lab and the volume of the 
jar to the water level, you can calculate how much gas you produced. (Plug P, V, T, and R into PV = nRT, 



387 



www.ckl2.org 



and solve for n.) It turns out, however, that you must make a correction before you plug in the pressure 
value. Since the collecting jar is a closed container and it has liquid water in the bottom of it, then it 
will contain the vapor pressure of water at this temperature. So the outside pressure in the lab tells you 
the pressure inside the collecting jar but it doesn't tell you how much of that pressure is due to the gas 
collected and how much is due to water vapor. You must get a table of the vapor pressure of water at each 
temperature and look up the vapor pressure of water at the temperature of your lab and then subtract 
that pressure from the total pressure in the collecting jar. The result will be the actual pressure of the gas 
collected. 

Example 

Some hydrogen gas was collected over water in the lab on a day that the atmospheric pressure was 755 torr 
and the lab temperature was 20°C Hydrogen gas was collected in the collecting jar until the water levels 
inside and outside the jar was equal. What was the partial pressure of the hydrogen in the collecting jar? 

Solution 

The total pressure in the collecting jar is 755 torr and is equal to the sum of the partial pressure of hydrogen 
in the jar and the vapor pressure of water at 20°C From the table, the vapor pressure of water at 20°C is 
17.5 torr. 

Partial pressure of H2 = 755 torr - 17.5 torr = 737 torr 

Lesson Summary 

• Molecules of liquid may evaporate from the surface of a liquid. 

• When molecules of a liquid evaporate, the remaining liquid cools. 

• Gas molecules in contact with their liquid may condense to liquid form. 

• If a liquid is placed in a closed container, eventually vapor pressure equilibrium will be reached. 

Review Questions 

1. A flask half-filled with water is sealed with a stopper. The space above the water contains hydrogen 
gas and water vapor in vapor pressure equilibrium with the liquid water. The total pressure of the 
two gases is 780. mm of Hg at 20. °C. The vapor pressure of water at 20. °C is 19 mm of Hg. What is 
the partial pressure of the hydrogen gas in the flask? 

2. Describe all the reasons that the remaining liquid cools as evaporation occurs. 

3. Describe all the reasons that the remaining gas gets hotter as condensation occurs. 

4. The apparatus above can be used to determine the vapor pressure of benzene. With a vacuum in the 
top of the tube, the mercury rises to the height shown. When a small amount of liquid benzene is 
injected into the space at the top of the tube, it floats on the mercury. The benzene will evaporate 
and eventually reach vapor pressure equilibrium. The mercury in the tube will pushed down further 
by the pressure of the benzene vapor in the tube. Neglecting the effect of the liquid benzene, what 
would be the calculated vapor pressure of benzene? 

5. Water vapor and hydrogen gas are sealed in a cylinder fitted with a piston at 60°C The partial 
pressure of the hydrogen gas is 0.35 atm and the vapor pressure of the water is 0.20 atm at this 
temperature. The total pressure in the cylinder is 0.55 atm. If the piston is pushed down until the 
volume is half the original volume, what will be the pressure in the cylinder? 

www.ckl2.org 388 



Vocabulary 

condensation The process whereby a gas or vapor is changed to a liquid. 

equilibrium vapor pressure The pressure that is exerted, at a given temperature, by the vapor of a 
solid or liquid in equilibrium with the vapor. 

evaporation The escape of molecules from a liquid into the gaseous state at a temperature below the 
boiling point. 

heat of condensation The quantity of heat released when a unit mass of a vapor condenses to liquid 
at constant temperature. 

heat of vaporization The quantity of heat required to vaporize a unit mass of liquid at constant tem- 
perature. 

vapor The gaseous phase of a substance that exists even though the temperature is below the boiling 
point of the substance. 



13.4 Boiling Point 

Lesson Objectives 



• The student will know the relationship between boiling point, vapor pressure, and ambient pressure. 

• Given a vapor pressure table for water, and the ambient pressure, the student will be able to determine 
the boiling point of water for those conditions. 

Introduction 

If you want hard-boiled eggs at home, you can probably put the eggs in boiling water for about eight 
minutes to accomplish it. If you go camping in the Rocky Mountains at an altitude of 10,000 feet, you 
will find that an egg placed in boiling water for eight minutes is not hard boiled. In fact, even after twelve 
minutes in boiling water, the egg may still be a little too runny for your tastes. 

Normal Boiling Point 

Imagine you are boiling water in a place where the atmospheric pressure is 1.00 atm. In the boiling water, 
a large bubble forms near the surface of the liquid water and remaining at the same size rises to the top of 
the water and the gas escapes into the air. If the pressure of the gas inside that bubble had been less than 
1.00 atm, the outside pressure of the atmosphere would have crushed the bubble and it would not have 
existed. If the pressure of the gas inside that bubble had been greater than 1.00 atm, the bubble would 
have expanded to a larger size instead of remaining at the same size. The fact that the bubble remained at 
the same size indicates that the gas pressure inside that bubble was the same as the atmospheric pressure. 

When you are heating water in an effort to boil it, gas bubbles cannot form until the water can produce 
a vapor pressure equal to the surrounding air pressure. The hotter the water gets, the higher its vapor 
pressure becomes but only when that vapor pressure equals the surrounding atmospheric pressure can the 

389 www.ckl2.org 



water form bubbles in the process we call boiling. A liquid cannot boil until its vapor pressure is equal to 
the pressure on the surface of the liquid. The actual definition of boiling point is that temperature at 
which the vapor pressure of the liquid equals the surrounding pressure. 

If you are measuring boiling points at the normal sea level atmospheric pressure of 1.00 atm, a liquid more 
volatile than water such as chloroform will boil at 61.3°C. This is because the vapor pressure of chloroform 
is 1.00 atm at 61.3°C. The vapor pressure of ethanol reaches 1.00 atm at a temperature of 78.4°C and 
therefore, that is the normal boiling point of ethanol. 

Boiling Points Change with Changes in Pressure 

Since liquids boil when their vapor pressure becomes equal to surrounding pressure, then if the surrounding 
pressure is lower, liquids will boil at lower temperatures. At higher altitudes, atmospheric pressure is lower. 
In cities whose altitude is around 5,000 feet, water boils at 95°C instead of at 100°C and at 10,000 feet, 
water boils around 90° C. The water boils in normal fashion but its temperature is lower and therefore, 
cooking in boiling water takes a longer time. In situations where boiling is used to purify water or sterilize 
equipment, the lower temperature of boiling water requires concern. 

If a container of water is placed in a bell jar and a vacuum pump attached so that the air pressure around 
the water can be greatly reduced, water may be made to boil at very low temperatures. (See Figure 
13.11). 



To Vacuum Pump 




Water Temperature = 20 C 

Figure 13.11: A beaker of water under a bell jar with lowered pressure. 

At room temperature, 20°C, the vapor pressure of water is 17.5 mm of Hg so if the pressure in the bell 
jar is reduced to 17.5 mm of Hg, water will boil at 20° C. The appearance of the boiling water is the same 
as it is at 100°C, with steam coming off and so on, but the water can be removed from the bell jar and 
poured on your hand and the temperature is only 20° C. When you look up the boiling point of a liquid, 
the reference will be to the normal boiling point which means the boiling point when the surrounding 
pressure is 1.00 atm. 

If the surrounding pressure is less than 1.00 atm, the boiling points of liquids will be lower. Conversely, if 
the surrounding pressure is greater than 1.00 atm, the boiling points of liquids will be higher. It's fairly 
unusual to find atmospheric pressures greater than 1.00 atm except during storms, but it's easy enough to 
raise the surrounding pressure in a laboratory situation. If we use a strong container with a lid that screws 
on very tight, we can boil water in the container and as water vaporizes and the temperature of both the 
air and the water vapor increase, the gas pressure in the container will increase. As the pressure in the 
container increases, the boiling point of the water increases. The vapor pressure of liquid water at 120°C 
is 2.0 atm. Therefore, if we can raise the pressure inside a sealed container to 2.0 atm, water will not boil 
in the container until its temperature is 120°C This is the concept that is used in pressure cookers and 
rice cookers. The cooking pot has a lid that can be sealed tightly and a valve in the lid that will open 
slightly when the pressure inside the container reaches 2.0 atm. Water and whatever food you wish to 

www.ckl2.org 390 



cook is placed inside the pressure cooker and it is set on the stove. The pressure and therefore the boiling 
point of water increases inside the container until the pressure reaches 2.0 atm. If the pressure goes beyond 
2.0 atm, the little valve opens and lets out some gas so that the pressure remains at 2.0 atm. The valve 
can be opened and closed any number of times to keep the inside pressure at 2.0 atm. The temperature of 
the boiling water inside will be 120°C under these conditions and the food will cook in as little as one-third 
the normal time. 



Lesson Summary 

• The boiling point of a liquid is the temperature at which the vapor pressure of the liquid becomes 
equal to the surrounding pressure. 

• The normal boiling of a liquid is the temperature at which the vapor pressure of the liquid becomes 
equal to 1.00 atmosphere. 



Review Questions 



1. What happens to the boiling point of a liquid if the pressure exerted on the surface of the liquid is 
increased? 

2. How can you make water boil without heating it? 

3. Fill in the diagram with either "high" or "low" to show how intermolecular forces of attraction influence 
the volatility, vapor pressure, and boiling point of a substance. 



weak 




Wlipti TVTF arp 


strona 




1 




i 




i 






f 


volatility is 




volatility is 



I 



vapor pressure is 



boiling point is 



i 



vapor pressure is 



boiling point is 



Vocabulary 

boiling point The temperature at which the vapor pressure of a liquid equals the surrounding pressure. 



normal boiling point The temperature at which the vapor pressure of a liquid equals 1.00 atmosphere. 

391 www.ckl2.org 



13.5 Heat of Vaporization 

Lesson Objectives 



• The student will be able to calculate energy changes during phase changes. 

• The student will be able to explain the slopes of various parts of heating and cooling curves. 

Introduction 

In order to vaporize a liquid, heat must be added to raise the kinetic energy (temperature) to the phase 
change temperature and then more heat must be added to provide the necessary potential energy to 
separate the molecules in the liquid form to the gaseous form. 

The Potential Energy Stored in Gases 

The difference between the liquid phase and the gas phase of a substance is essentially the distance between 
the molecules. Since the molecules attract each other and they are separated by a greater distance in the 
gaseous phase than in the liquid phase, the molecules in gaseous phase possess more potential energy 
than in the liquid phase. When a substance changes from the liquid phase to the gaseous phase, work 
(the physics kind of work as in force times distance) must be done on the molecules to pull them away 
from each other. The work done separating the molecules is stored in the molecular structure as potential 
energy. If the molecules are allowed to go back together as in the liquid phase, the potential energy is 
released - exactly the same amount that was needed to separate the molecules. This potential energy 
stored in molecules in the gaseous phase is called the heat of vaporization. The heat of vaporization 
(a/Zvap) f° r water is 540 calories/gram (2.26 kJ/g) at the normal boiling point. Because of the strength 
of the polar attractions holding water molecules together in the liquid form, water has a fairly high heat 
of vaporization. Ammonia, NH3, and ethanol, C2H5OH, which are also polar molecules have heats of 
vaporization of 1.38 kJ/g and 0.84 kJ/g respectively. 

Example 

How much heat in kJ is necessary to vaporize 100. grams of ammonia at its boiling point? 

Solution 

Heat, Q = (mass)(A// V Ap) = (100. g)(1.38 kJ/g) = 138 kJ 

The boiling point of ammonia is — 33°C. It is very important to understand that the ammonia is at the 
boiling point before the heat of vaporization is added and after the heat of vaporization is added, the 
ammonia is STILL at the boiling point. All the energy involved in the heat of vaporization is absorbed 
by the substance as potential energy - none of it goes into kinetic energy and therefore, the temperature 
cannot change. A common question asked of students to determine if they understand this point is to ask 
which would produce a more severe burn, spilling boiling water at 100°C on your skin or being burned 
by gaseous water at 100°C? It may seem at first that since they are both at the same temperature, they 
would do the same damage, but in fact, the gaseous water would first deliver a tremendous amount of heat 
to your skin as the gas condensed to water (giving up the heat of vaporization) and after you had a burn 
from that, then you would have 100°C boiling water on the skin that was already burned. 

www.ckl2.org 392 



Heating and Cooling Curves 

The addition of heat before, during, and after a phase transition can be analyzed with the help of a heating 
curve. In the heating curve below, a sample of water at 20°C and 1.00 atm pressure had equal quantities 
of heat added to it per unit time. 



[30- 


t / 


iao- 


/ 


ItG- 


/ 


IOG 


/ 


VO- 




BO- 




70- 




feO- 






/ 


-■; 


- Jr 


30- 




ZO- 




to- 

■ ! 





Hear Added 



Figure 13.12: The heating curve for water at normal pressure. 



Between the temperatures of 20°C and 100°C, all the heat added was absorbed as kinetic energy and 
therefore the temperature increased. Once the water reached the boiling point, even though the same 
amount of heat was added per unit time, the temperature did not increase. Thus, this energy added to the 
water DID NOT become kinetic energy. All the heat added to the sample during the time the slope of the 
line is zero went into potential energy. This energy represents the heat of vaporization and was used to 
do the work of separating the liquid molecules into the gaseous form (greater distance between molecules). 
During this flat line period, an observer would see that the water was changing into gas but all of it, both 
the part in liquid form and the part in gaseous form, would be at exactly 100°C The temperature cannot 
increase until all the liquid has been converted to gas. At 1.00 atm pressure, it is impossible to get water 
hotter than 100°C No matter how much heat you add to it, all that happens is that the water boils faster 
and converts to gas faster but its temperature will never exceed 100°C It is sort of like the water decides 
where the heat goes, and the water decides that ALL the heat added will be used to change phase and 
none of it will be used to raise temperature. Once all the water is in the gaseous form, the heat once 
again becomes kinetic energy and the temperature of the gas rises. When the gas is cooled, it follows this 
same curve in reverse. As it is cooled, the same flat line appears while the heat of vaporization (heat of 
condensation now) is removed, and then the temperature may go down again. 



Specific Heat 

Thermodynamic data (melting point, boiling point, heat of melting, heat of vaporization) for almost 
all elements and thousands of compounds are determined by laboratory activity and placed in reference 
books. Much of this thermodynamic data is also available now on the internet. Another piece of useful 
thermodynamic data about substances is called specific heat. The specific heat for a substance is the 
amount of heat required to raise 1.00 gram of the substance by 1.00°C. The symbol C is used for specific 
heat and the value for liquid water is 4.18 J/g°C. 

393 www.ckl2.org 



Example 

How much heat is required to raise the temperature of 25 grams of water from 15°C to 55°C? 

Solution 

Q = (mass) (C) (At) = (25 g)(4.18 J/g.°C)(40.°C) = 4180 J = 4.18 kJ 

Example 

How much heat is required to raise 25 g of water from 25°C to gaseous water at 100°C? 

Solution 

In this problem, you have to calculate the heat to raise the temperature from 25°C to 100°C and then 
calculate the heat of vaporization for the 25 g of water. 

Heat raising temp = (mass) (specific heat)(A?) = (25 g)(4.18 J/g.°C)(75°C) = 7840 J = 7.48 kJ 
Heat va p 0r i z i n g = (mass)(heat of vaporization) = (25 g)(2.26 kJ/g) = 57 kJ 
Total heat = 7.48 kJ + 57 kJ = 64 kJ 

Lesson Summary 

• The heat of vaporization of a liquid is the quantity of heat required to vaporize a unit mass of that 
liquid at constant temperature. 

• The energy released when a gas condenses to a liquid is called the heat of condensation. 

• The specific heat of a substance is the amount of heat required to raise the temperature of one gram 
of the substance by 1.0°C. 

Review Questions 

1. How much heat is required to vaporize 200. grams of water at 100. °C and 1.00 atm pressure? a//vap 
for water is 2.25 kJ/g. 

2. How much heat is required to raise 80.0 grams of water from 0°C to 100. °C with no phase change 
occurring? The specific heat of water is 4.18 J/g.°C 

Further Reading / Supplemental Links 

• Chemistry, Matter and Its Changes, Fourth Edition, Chapter 12: Intermolecular Attractions and the 
Properties of Liquids and Solids, James E. Brady and Fred Senese, John Wiley & Sons, Inc. 2004. 
http : //www . sparknotes . com/testprep/books/sat2/chemistry/chapter5sect ion5 . rhtml 

• http : //www . kentchemistry . com/links/matter/heatingcurve . htm 

Vocabulary 

heat of condensation The quantity of heat released when a unit mass of a vapor condenses to liquid 
at constant temperature. 

www.ckl2.org 394 



heat of vaporization The quantity of heat required to vaporize a unit mass of liquid at constant tem- 
perature. 



Image Sources 



(i 

(2 
(3 
(4 
(5 

(6 

(7 
(8 

(9 
(10 

(11 
(12 



Richard Parsons. . CC-BY-SA. 

Richard Parsons. . CC-BY-SA. 

Richard Parsons. . CC-BY-SA. 

Richard Parsons. . CC-BY-SA. 

Richard Parsons. The centers of positive and negative charge.. CC-BY-SA. 

Richard Parsons. Gas Collection Over Water.. CC-BY-SA. 

Richard Parsons. . CC-BY-SA. 

Richard Parsons. . CC-BY-SA. 

Richard Parsons. Standard Distribution Curve. CC-BY-SA. 

Richard Parsons. The Three Phases of Matter.. CC-BY-SA. 

Richard Parsons. The heating curve for water at normal pressure.. CC-BY-SA. 

Richard Parsons. . CC-BY-SA. 



395 www.ckl2.org 



Chapter 14 



The Solid State 



14.1 The Molecular Arrangement in Solids Con- 
trols Solid Characteristics 

Lesson Objectives 

• The student will describe the molecular arrangement of solids. 

• The student will use the molecular arrangement in solids to explain the incompressibility of solids. 

• The student will use the molecular arrangement in solids to explain the low rate of diffusion in solids. 

• The student will use the molecular arrangement in solids to explain the ability of solids to maintain 
their shape and volume. 



Introduction 

There are many ways to classify solids but the broadest categories are crystalline solids, those with a 
highly regular arrangement of their particles, and amorphous solids, those with considerable disorder 
in their structures. The regular arrangement of the particles in a crystalline solid produces the beautiful, 
characteristic shapes of crystals. These structures are represented by a crystal lattice, a three-dimensional 
system of points designating the positions of the component particles (atoms, ions, or molecules). There 
are also many important amorphous solids. An example is glass, which is best represented as a solid where 
the components were frozen in place before they could get into the organized structure of a crystal. Because 
of the disorder in glass, some chemists have referred to glass as a super-cooled liquid. 



The Molecular Arrangement in Solids 

The molecular arrangement in a solid is one where the molecules (or atoms, or ions) are held in a tightly 
packed, highly organized pattern, as in the below figure. 

The molecular arrangement in a solid. (Source: Richard Parsons. CC-BY-SA) 
www.ckl2.org 396 




II I 



In a solid, the intermolecular forces of attraction have completely overcome molecular motion and the 
movement of the particles has been reduced to vibration in place. There are only tiny spaces between the 
molecules, not nearly enough space for the particles to move past one another. 

The Characteristics of Solids 

The intermolecular forces of attraction in solids hold the particles so tightly in place that they cannot 
pull away from each other to expand their volume nor can they flow past one another to change shape. 
Therefore, solids hold their own shape and volume regardless of their container. There is very little empty 
space in the solid structure so solids are virtually incompressible and since molecules can't pass each other 
in the structure, diffusion is essentially non-existent beyond the surface layer. 

Lesson Summary 

• Molecules in a solid maintain both their own shape and their own volume. 

• Solids are virtually incompressible and have little diffusion beyond the surface layer. 

• The molecular arrangement in solids is a highly organized, tightly-packed pattern with small spaces 
and molecular motion reduced to vibration in place. 

Review Questions 



. Fill 


in the types of phase 


changes left blank in 


the chart below. 






Solid - 


» Liquid 








Liquid 


— > Gas 








Solid - 


» Gas 


Sublimation 






Liquid 


-> Solid 








Gas — » 


Liquid 








Gas — » 


Solid 


Deposition 



14.2 Melting 

Lesson Objectives 



The student will explain why it is necessary for a solid to absorb heat during melting even though 
no temperature change is occurring. 

Given appropriate thermodynamic data, the student will calculate the heat required to raise temper- 
atures of a given substance with no phase change. 

Given appropriate thermodynamic data, the student will calculate the heat required to melt specific 
samples of solids with no temperature change. 

Given appropriate thermodynamic data, the student will calculate the heat required to produce both 
a phase change and temperature change for a given sample of solid. 

397 www.ckl2.org 



Introduction 

The melting point of a substance, like its boiling point, is directly related to the strength of the forces 
of attraction between molecules. Low melting points are typical of substances whose forces of attraction 
are very weak, such as hydrogen gas whose melting point is — 259°C. High melting points are associated 
with substances whose forces of attraction are very strong, such as elemental carbon whose melting point 
is greater than 3500° C. 

Melting Points 

Solids, like liquids, have vapor pressure. The vapor pressure of a solid is generally very low because the 
forces of attraction in solids dominate molecular motion. The vapor pressure of solids, also like liquids, 
increases with temperature. The vapor pressure of liquid water is 760 mm of Hg at 100°C and decreases 
as the temperature decreases (non-linearly) to 4.6 mm of Hg at 0°C Solid water (ice) has very low vapor 
pressures because of both strong forces holding the molecules together and the fact that our interaction 
with ice is usually at low temperatures. For example, at -83°C, the vapor pressure of ice is 0.00025 mm 
of Hg. As ice is heated, its vapor pressure increases. At 0°C, the vapor pressure of ice is 4.6 mm of Hg 
which also happens to be the vapor pressure of water at 0°C. In fact, for all substances, their solids and 
liquids have the same vapor pressure at the melting point. The melting point of a solid is defined as the 
temperature at which the vapor pressure of the solid and liquid are the same. 

Heat of Fusion 

The phase change from solid to liquid, melting, has many similarities to vaporization. The solid must reach 
its melting point before the molecules can enter the liquid phase. The molecules in liquid phase, however, 
are farther apart than the molecules in the solid phase. Since the molecules attract each other, increasing 
the distance between them (from solid structure to liquid structure) requires work. Force must be exerted 
to pull the molecules away from each other. The work done in separating the molecules is stored in the 
molecules as potential energy in the liquid phase. This is the same process that occurs when the heat of 
vaporization must be added to liquid molecules to get them into the gaseous phase. In the case of melting, 
this potential energy is called the heat of fusion or the heat of melting. 

The word 'fusion' is used several times in science with different meanings. You need to note the context 
of the use of the word to know which meaning is intended. In this case, fusion is the name of the liquid to 
solid phase change. When a solid melts, the heat of fusion must be added and when a liquid freezes (fuses) 
back to solid, the heat of fusion is given off. The heat of fusion for water is 334 Joules/gram. 

Example 1 

How much heat must be added to 25 grams of ice at 0°C to convert it to liquid water at 0°C? 
Solution 

Q = (mass)(A// FUS ioN) = (25 g)(334 J/g) = 8350 J = 8.4 kJ 

Heating Curves 

Phase changes are often analyzed with the help of a heating curve. The heating curve for water at 1.00 atm 
pressure appears in Figure 14.1. Many substances will have heating curves similar to the one for water. 
The differences will be in the length of flat lines and the slopes of the inclined lines. 

www.ckl2.org 398 



The Heating Curve for Water at 1 .00 atm 



120- 

i-ij- 
[03 " 








1 


90- 






* 




80- 










o^ 70- 

| 60- 

1 - 




A 






30- 










20- 










10- 


2 / 








.»- 


/ 





Figure 14.1: The heating curve for water. 

The section of the curve labeled "1" represents the time the sample is completely in the form of ice and is 
being heated. All the heat added is going into kinetic energy and so the temperature of the ice is increasing. 
The slope of the line is related to the specific heat of ice - the amount of heat required to raise 1.00 gram 
of ice by 1.00°C. When the sample reaches 0°C, that is, the melting point, all the heat added starts going 
into potential energy. During the section of the graph labeled "2," the sample is being converted from 
solid to liquid. Both solid and liquid are present at this time, but all of the sample, both solid and liquid 
are at 0°C. No heat will go into kinetic energy until the entire sample is converted to liquid. Therefore, 
the temperature remains constant during this period. The total amount of heat added during section "2" 
would represent the heat of melting for this sample. Once the entire sample has been converted to liquid, 
the added heat would once again go into kinetic energy and the temperature would increase. The slope of 
the line in section "3" is related to the specific heat of liquid water. The temperature continues to rise as 
heat is added until the boiling point is reached. The temperature again remains constant through section 
"4" and the added heat becomes potential energy providing the heat of vaporization. During section "4," 
both liquid water and gaseous water are present. Once the entire sample has been converted to gas, the 
added heat again becomes kinetic energy and the temperature of the gas increases. Gaseous water also 
has its own specific heat. 

Example 2 

A 2, 000. gram mass of water in a calorimeter has its temperature raised by 3.00°C while an exothermic 
chemical reaction is occurring. How much heat, in joules, is transferred to the water by the heat of reaction? 

Solution 

The heat is calculated by determining the heat absorbed by the water. This amount of heat is the product 
of three factors, 1) the mass of the water, 2) the specific heat of water, and 3) the change in temperature 
of the water. 

heat = (mass of water) (C water )(At) heat = (2000.g)(4.18 J/g -° C)(3.00°C) = 25,080 joules = 25, 100 joules 

Example 3 

A 1,000. gram mass of water whose temperature was 50. °C lost 33,440 joules of heat over a 5-minute 
period. What was the temperature of the water after the heat loss? 

Solution 

heat = (mass) (C w ) (At) 

heat -33,400 ioules 

Af = 7T- = -, b — ; r = -8.00°C 

mass x C w (1000. g)(4.18 J/g -° C) 

399 www.ckl2.org 



If the original temperature was 50. °C and the temperature decreased by 8°C, then the final temperature 
would be 42°C. 

Example 4 

Use the thermodynamic data given to calculate the total amount of energy necessary to raise 25 grams of 
ice at -20°C to gaseous water at 120°C. 



Melting point of ice = 0°C 
a//vap for water = 2260 J/g 



Boiling point of water = 100°C 
A//pusiON fo r water = 334 J/g 



C ice = 2.11 J/g.°C 



C water = 4.18 J/g.°C 



C„ater vapor = 1.84 J/g.°C 



Solution 

There will be five steps in the solution process. 

1. We must calculate the heat required to raise the temperature of the ice from -20°C to 0°C 

2. We must calculate the heat required to provide the heat of melting - to change ice at 0°C to water 
at 0°C. 

3. We must calculate the heat required to raise the temperature of the liquid water from 0°C to 100°C 

4. We must calculate the heat required to provide the heat of vaporization - to change liquid water at 
100°C to gaseous water at 100°C. 

5. Finally, we must calculate the heat required to raise the temperature of the gaseous water from 100°C 
to 120°C. 

Raising the temperature of ice, Q = mCAt = (25 g)(2.11 J/g.°C)(20°C) = 1051 J 

Melting ice to liquid, Q = (mass)(A// FUSI o N ) = (25 g)(334 J/g) = 8350 J 

Raising the temperature of liquid, Q = mCAf = (25 g)(4.18 J/g.°C)(100°C) = 10450 J 

Vaporizing liquid to gas, Q = (mass)(A//vAp) = (25 g)(2260 J/g) = 56500 J 

Raising the temperature of gas, Q = mCAt = (25 g)(1.84 J/g.°C)(20°C) = 920 J 

The sum of all five steps is 77, 000 J = 77 kJ 

The cooling process would be exact reverse of the heating curve. If water in the gaseous phase is cooled, 
each 1.84 Joules of heat removed would lower the temperature of 1.00 g of gas by 1.00°C. When the gaseous 
water reaches the boiling point (also the condensation point), each gram of gaseous water that condenses 
to liquid will release 2260 Joules of heat. Once all the water is in the liquid form, the removal of each 
4.18 Joules of heat by cooling will cause the temperature of 1.00 g of water to cool by 1.00°C At the 
freezing point (also the melting point) 334 Joules of heat must be removed to convert each gram of liquid 
water to ice. When the entire sample of water is in the form of ice, 2.26 Joules of heat must be removed 
to cool each gram by 1.00°C. For all phase changes, going up in temperature or down in temperature, the 
entire sample will change phase before any temperature change will occur. 

Since both melting points and heats of fusion are dependent on the strength of the attractive forces between 
molecules, a solid with a low melting point will usually also have a low heat of fusion and a solid with a 
high melting point will have a high heat of fusion. 



Lesson Summary 

• Solids melt when the vapor pressure of the solid equals the vapor pressure of the liquid. 
www.ckl2.org 400 



• Heat must be absorbed by a solid to become a liquid even though the temperature remains the same. 
The quantity of heat absorbed per unit mass is called the heat of fusion. 

• Stronger forces of attraction between particles in solids produce higher melting points and higher 
heats of fusion. 

• During a phase change, all the added energy goes into the heat of fusion (potential energy) and none 
goes to kinetic energy (raising temperature). 

Review Questions 

Use the thermodynamic data given in the Table 14.1 to answer problems 1-5. 

Table 14.1: Thermodynamic Data of Various Substances 



Water 



Cesium, Cs 



Silver, Ag 



Melting Point 
Boiling Point 

Affusion 



o°c 

100.°C 
334 J/g 

^//vaporization 2260 J/g 

Specific Heat, C, for 2.01 J/g ° C 

Gas 

Specific Heat, C, for 4.18 J/g ° C 

Liquid 

Specific Heat, C, for 2.09 J/g ° C 

Solid 



29°C 
690.°C 
16.3 J/g 
669 J/g 
0.167 J/g °C 

0.209 J/g -° C 

0.251 J/g-°C 



962°C 

2162°C 
105 J/g 
2362 J/g 
0.159 J/g °C 

0.294 J/g -° C 

0.235 J/g -° C 



1. How many Joules are required to melt 100. grams of silver at its normal melting point with no 
temperature change? 

2. How many Joules are required to boil 150. grams of cesium at its normal boiling point with no 
temperature change? 

3. How many Joules are required to heat 200. g of liquid water from 25°C to steam at 125°C under 
normal pressure? 

4. How many Joules are required raise the temperature of 1.00 gram of water from -269°C (the current 
temperature of space) to 1.60 x 10 15 °C (the estimated temperature of space immediately after the 
big bang)? 

5. How many Joules are required to raise the temperature of 1000. g of cesium from -200. °C to +200. °C? 

6. Why does the boiling point of water increase with increasing surrounding pressure? 

7. Why must heat be absorbed to melt a solid even though both the solid and the liquid are at the 
same temperature? 



Heating Curve for Water at 1.00 atm Pressure 




401 



www.ckl2.org 



8. What is happening to the water in section B? 

9. What is happening to the water in section A? 

10. Why are the slopes of the lines in sections A, C, and E different? 

Vocabulary 

crystal A solid consisting of plane faces and having definite shape with the atoms arranged in a repeating 
pattern. 

freezing The phase change from liquid to solid. 

freezing point The temperature at which a liquid changes to a solid. 

fusion 

1. The change of a liquid to a solid. 

2. A nuclear reaction in which two or more smaller nuclei combine to form a single nucleus. 

heat of condensation The quantity of heat released when a unit mass of vapor condenses to a liquid 
at constant temperature. 

heat of fusion The quantity of heat released when a unit mass of liquid freezes to a solid at a constant 
temperature. 

heat of vaporization The quantity of heat absorbed when a unit mass of liquid vaporizes to a gas at 
constant temperature. 

joule A basic unit of energy in the SI system, equal to one Newton-meter. 

melting The phase change from solid to liquid. 

melting point The temperature at which a substance changes from the solid phase to the liquid phase. 

14.3 Types of Forces of Attraction for Solids. 

Lesson Objectives 

• The student describe the metallic bond and explain some of the solid characteristics that are due to 
metallic bonding. 

• Given the characteristics of a solid such as conductivity of solid and liquid phase, solubility in water, 
malleability, and so on, the student will be able to identify the type of solid, i.e. the attractive forces 
holding the solid in the solid form. 

Introduction 

The range of melting points for the various types of solids is extremely wide. Substances with very weak 
London dispersion forces like helium, will melt at only a couple of degrees above absolute zero whereas 
solid substances like asbestos and diamond do not melt until the temperature is in excess of 3500°C 

www.ckl2.org 402 



Solids Held Together by Intermolecular Forces of Attraction 

Each of the intermolecular forces of attraction discussed in the chapter on liquids (London dispersion 
forces, polar attractions, and hydrogen bonds) will also produce solids if the temperatures are low enough. 
Those intermolecular forces of attraction have the same causes in solids as they do in liquids but because 
of the lower temperatures and the closeness of the molecules in solids, the forces will be more effective in 
pulling the molecules together. The solids formed due to these intermolecular forces of attraction will be 
crystalline solids. 

Solids held together by London dispersion forces are not soluble in water, nor are they good conductors 
of electricity or malleable. In terms of solubility, the general rule is "like dissolves like." This means that 
polar solvents dissolve polar or ionic substances but not non-polar and non-polar solvents dissolve non- 
polar solutes but not polar. Since water is a strongly polar molecule, it will dissolve most polar solids. 
Polar solids are not good conductors and are not malleable. 

Ionic Solids 

Ionics solids are held together by the electrostatic attraction between oppositely charged ions. The ions 
are formed into various types of crystal lattice structures depending on the comparative sizes of the ions 
and the charges on each. These ionic charges are full 1+, 2+, 1-, 2-, and so on, charges and so they 
are considerably stronger than either polar attractions or the especially strong hydrogen bonds. This will 
cause the melting points for ionic substances to be quite high compared to the substances we have been 
considering. For example, the melting point of sodium chloride, NaCl, is 801°C and the melting point of 
calcium sulfate, CaSO^, is 1460° C. 

In solid state, the ions in ionic solids are held firmly in position and there are no spaces large enough for 
the ions to move through even if they could escape the forces of attraction. Since the ions cannot move, 
ionic solids are non-conductors of electricity. When the solid is melted to a liquid, however, the ions are 
free to migrate and therefore, ionic liquids are good conductors of electric current. 

In ionic crystals, the positive and negative ions occupy positions so the ions will attract the maximum 
possible number of oppositely charged ions. (See Figure 14.2). 




Figure 14.2: A crystal. 

In the NaCl lattice, each sodium ion is equally attracted to six chloride ions. In the sketch you can see 
each sodium surrounded by four chlorides, and in a three-dimensional structure, there would be another 
chloride above the sodium ions in the layer above and another one in the layer below, for a total of six. 



403 



www.ckl2.org 



The word malleable (in science) refers to a substance that can be pounded or beaten into another shape 
without breaking. Ionic solids are NOT malleable. (See Figure 14.3). 




Figure 14.3: Ionic solids are not malleable. 



If an ionic crystal is struck, the ions are driven down a layer so that they will be next to like charges. 
These charges will repel and the crystal will shatter. 

Ionic solids are usually quite soluble in water and the water solutions of ionic solids are good conductors 
of electricity because of the freedom of the ions to migrate through the solution. 



Metallic Solids 



Of the 81 elements which can be clearly classified as metals, all of them except mercury are solids at 
room temperature. Any model to explain the bonding in metallic solids must account for the properties 
of metals, some of which are quite unusual. The properties of metals include 1) excellent conductors of 
heat and electricity in both solid and liquid phase, 2) malleable, 3) most of them are white and shiny, 4) 
metallic solids are not soluble in any common solvent, polar or non-polar, and 5) they have a wide-range 
of melting points mostly higher than the melting points of polar solids. 

The simplest bonding model that has been proposed to explain metallic behavior is the metallic bond, 
which envisions a regular pattern of cations surrounded by a "sea of electrons." The metal ions (all of 
the metal atom except the valence electrons) occupy positions in a lattice structure while the mobile, 
free-moving, sea of valence electrons occupy all of the overlapping valence shell area. See below figure. 

The metallic bond consists of non-directed bonds in which a "sea of electrons" surrounds all the bonded 
atoms. (Source: Richard Parsons. CC-BY-SA) 



www.ckl2.org 



404 




All the atoms are bonded in a single bond that includes the entire piece of metal. The bond is non- 
directional and does not hold adjacent ions in position relative to each other. 

For purposes of comparison, consider the covalent bonding in a trigonal planar molecule below. The central 
atom in this molecule contains three pairs of electrons. The electrons take positions as far away from each 
other as possible due to electrostatic repulsion. Therefore, the pairs of electrons maintain directed positions 
at angles of 120° from each other. The atoms that share these electron pairs in the covalent bond must be 
placed so that the shared electrons are in the overlapped orbitals of both atoms. Therefore, these bonded 
atoms may not move with respect to each other. The atoms hold their relative positions because of the 
directional bonds. Neither the bonding electrons nor the atoms are free to move with respect to each 
others. 

Directional bonding in trigonal planar molecule. (Source: CK-12 Foundation. CC-BY-SA) 



a: 




The model of the metallic bond, however, provides for mobile electrons that are free to move throughout 
the entire piece of metal and therefore, provides the means for the metal to be an excellent conductor of 
heat and electricity. Extra electrons pushed onto the metal on one side can easily move through the valence 
electron shells to the other side. The metal ions are not directionally bonded to their immediate neighbors 
and this allows them to be pushed to new positions without breaking the bond. As long as an atom or ion 
is not separated from the piece of metal, its position can be significantly changed while it remains bonded 
to the other atoms. This malleability allows metal cubes to be pounded into flat sheets without breaking 
the bond. 

The freedom of the electrons on the surface of a piece of metal also allows the metal to absorb and emit 
many frequencies of light which accounts for the white, shiny appearance of many metals. The metals 
on the far left of the periodic table have the fewest valence electrons in the valence shells so the valence 
electrons in these metals would be least crowded, have the most freedom, and present the most complete 
metal character. The metals of families IA and IIA are excellent conductors, exceptionally malleable (soft 
enough to be cut with a spoon), and white and shiny in color. 

Other elements can be introduced into a metallic crystal relatively easily to produce substances known as 
alloys. An alloy is defined as a substance that contains a mixture of elements and has metallic properties. 
A fairly well-known alloy is brass which is an alloy composed of about two-thirds copper atoms and one- 
third zinc atoms. Sterling silver is an alloy composed of about 93% silver and 7% copper. 

405 www.ckl2.org 



Iron is a metal that is commonly alloyed with carbon to produce steel. The carbon forms directional 
bonds with some of the iron atoms to make steel less malleable than pure iron. Steel with less than 0.2% 
carbon remains somewhat malleable and is used for nails and cables. Steel with around 0.6% carbon is 
harder and is used for railroad rails and structural steel beams. Steel with around 1.5% carbon is very 
hard and is used for tools and cutlery. 

Malleability, ductility, and conductivity are examined, along with methods for extracting metals from ores 
and blending alloys. Metals (http: //www. learner. org/vod/vod_window.html?pid=811) 

Network Solids 

In some solids, all the atoms in the entire structure are bonded with covalent chemical bonds. These 
solids are a single giant molecule and are called network solids. When considering the strength of the 
bonds/attractions that hold various particles together in the solid state, the strongest of them all are 
covalent bonds. Therefore, network solids have the highest melting points of all solids. To melt a network 
solid requires enough molecular motion to disrupt covalent chemical bonds. Network solids are not soluble 
in any common solvent. Most network solids are non-conductors but graphite is an exception, it is a good 
conductor of electricity. Some examples of network solids are graphite, diamonds, mica, and asbestos. 

The solid structure in graphite involves large two-dimensional molecules of covalently bonded carbon atoms. 
The carbon atoms form flat sheets (like a sheet of paper) bonded in the fashion shown in Figure 14.4. 



Graphite: a two-dimensional network solid. 



I 



/N/N/N 

c c c 



e \,/ c \ e / e \ e /% 



Diamond: a three-dimensional network solid. 



\ c / C \c/ e \,/ < 



"<5 

c 



\</ c \ e /' 




Figure 14.4: One layer (sheet) of graphite, a dimensional network solid, and diamond, a dimensional 
network solid. 

Then layers of these sheets are laid on top of each other and the sheets are held together by much weaker 
London dispersion forces. The sheets are extremely strong in the two dimensions involving covalent bonds 
but the forces holding the sheets together are weak and easily broken. The flat sheets slide over each other 
readily and this makes graphite a good lubricant for metal parts. 

The mineral mica is also bonded in this two-dimensional network style. Mica is found in nature and appears 
as a rock but you can slide your fingernail between sheets and pull off large flats sheets of the rock. One 
type of mica, called muscovite mica, is transparent enough that you can see through several sheets. This 
material has been used to make small windows in furnaces so the operator can look in but the window is 
rock, not glass, so it won't melt. 



www.ckl2.org 



406 



Diamonds are giant molecules of carbon atoms bonded three-dimensionally in tetrahedral units. Every 
carbon atom in the structure is covalently bonded to four other carbon atoms. Diamonds are the hardest 
substance known and have one of the highest melting points of all substances. 

Some forms of asbestos are a one-dimensional network solid in which atoms are bonded in a chain. The 
result is a fibrous molecule that can be woven into fabric. (See Figure 14.5). 




Figure 14.5: Asbestos in the form of chrysotile. You can see fibers of asbestos in the upper left hand corner 
of the photo. 

Due to the high melting point, asbestos fabric was used to make heat resistant materials (fireman's gloves, 
furnace padding, clutch plates, etc.) for many years until it was determined that asbestos fibers are 
hazardous if inhaled. 

The following web site has data and explanatory reasons for the trends in melting and boiling points 
of some period 3 elements. Trends in Melting Points and Boiling Points for Period 3 elements (http: 
//www. creative-chemistry.org.uk/alevel/modulel/trends8.htm) 



Amorphous Solids 



Many important solids do not have the regular, repeating arrangement of atoms or molecules that is present 
in crystalline solids. Solids with irregular, unpredictable molecular organization are called amorphous 
solids. There are many solids that will form either crystalline or amorphous solids depending on how 
rapidly the liquid is cooled. Very rapid cooling of these substances frequently results in an amorphous 
solid whereas slow cooling produces crystalline solids. Amorphous solids have been described as appearing 
to have their molecules frozen in place before they had time to get into an organized pattern. Examples 
of amorphous solids are glass, paper, plastics, cement, and rubber. 

Amorphous solids are called solids because they maintain their shape and volume. Some researchers insist 
that certain amorphous solids will flow under pressure, which is a characteristic of liquids. Antique windows 
have been found that are thicker at the bottom than at the top. Some chemists claim this is because the 
glass very slowly, over a hundred years, flowed downward under gravitational force. Other chemists claim 
the antique glass being of different thicknesses was caused by flaws in the glass making process of a hundred 
years ago. There was no mention in the opinion of why the thicker part of the glass was always at the 
bottom and never at the top. 

Crystalline solids melt at sharply defined temperatures. Glass and some other amorphous solids, however, 
soften as they are heated. Some authors refer to amorphous solids as "not true solids", others call them 
"supercooled liquids," and still others insist that amorphous solids are absolutely solids. The importance 
of amorphous solids such as glass and plastics will insure that research on their structure continues. 



407 



www.ckl2.org 



Lesson Summary 

• One type of solid is formed by ionic solids in which the inter-particle forces of attraction are electro- 
static attractions due to the opposite charges of the ions. 

• One type of solid is formed by metallic atoms where a sea of electrons exerts a force of attraction on 
the positive ions (metallic bond). 

• Network solids have every atom in the structure attached to other atoms in the structure by covalent 
chemical bonds. 

• Amorphous solids are solids that cooled so rapidly, the molecules did not get into the tight, organized 
solid pattern. Due to their disorganized structure, amorphous solids have some properties more like 
liquids. 

Review Questions 

1. Identify the most important type of inter-particle force present in the following solids that is respon- 
sible for binding the particles into a solid. 



(a) He 


(b) NO 


(c) HF 


(d) BaCh 


(e) CH A 


(f) NaN0 3 


(g) co 2 


(h) CHCh 


(i) pure Mg 


(j) diamond 



2. Predict which substance in the following pairs would have the stronger force of attraction between 
molecules and justify your answer. 

(a) C0 2 or OCS 

(b) PF 3 or PF 5 

(c) Nal or I 2 

(d) H 2 or H 2 S 

(e) solid argon or solid sodium 

(f) HF or HBr 

3. In the following groups of substances, pick the one that has the requested property and justify your 
answer. 

(a) highest boiling point: HCl,Ar,F 2 

(b) highest melting point: H 2 0,NaCl,HF 

(c) lowest vapor pressure at 20°C: CI 2 , Br 2 , I 2 

4. An unknown solid is not soluble in water or CC/4. The solid conducts electricity and has a melting 
point of 800°C Identify the most likely attractive forces holding the particles in the solid state. 

5. An unknown solid is soluble in water but not in CC/4. The solid does not conduct electricity but 
its liquid does. The solid shatters when hammered and has a melting point of 1430°C. Identify the 
most likely attractive forces holding the particles in the solid state. 

6. Why would you expect ionic solids to have higher melting points that polar solids? 

7. Why does the melting point of water decrease with increasing surrounding pressure? 

8. List the following substances in order of increasing boiling points: BaCl 2 , H 2 , CO, HF, Ne, C0 2 . 

www.ckl2.org 408 



(a) H 2 

(b) CH-iOH 

(c) CH 2 Ch 

(d) KCl 

(e) CO 

Select your answers for questions 9, 10, and 11 from these choices. 

9. Which of these substance is most likely to be a solid at 25°C and 1.0 atm? 

10. Which of these substances is capable of hydrogen bonding? 

11. Which of the substances has its solid properties governed by London dispersion forces? 

12. Place these molecules, CF4, CaCh, and ICl, in order of decreasing melting points (highest first). 

(a) CFi > CaCh > ICl 

(b) CaCh > ICl > CF A 

(c) CaCh > CF 4 > ICl 

(d) ICl > CF 4 > CaCh 

(e) CF A > ICl > CaCh 

Further Reading / Supplementary Links 

• http : //learner . org/resources/series61 . html 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. The website has one 
video that relates to this lesson called Metals. 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 4-6 is on Intermolecular Forces. 

• Lesson 210 is on Heat Transfer Calculations. 

• http : //www . f ordhamprep . org/gcurran/sho/sho/lessons/lesson46 . htm 

• http : //dwb . unl . edu/teacher/nsf/cO 11 inks/www . ualberta . ca/ ! bderksen/f lor in . html 

• http : //www . sciencedaily . com/releases/2008/07/080704153507 . htm 

• http : //www . bestcrystals . com/crystals2 . html 

Vocabulary 

alloy A substance composed of a mixture of two or more elements and having metallic properties. 
conductivity The property of being able to transmit heat and/or electricity. 
conductor A substance that can transmit heat and/or electricity. 
ductility The property of a substance that allows it to be drawn into a wire. 
electrical conductivity The ability of a substance to transmit an electric current. 
malleable The property of being able to be hammered or rolled into sheets. 

409 www.ckl2.org 



metallic bond The attractive force that binds metal atoms together. It is due to the attractive force 
that the mobile electrons exert on the positive ions. 

specific heat The amount of energy necessary to raise 1.00 gram of a substance by 1.00°C. 

14.4 Phase Diagrams 

Lesson Objectives 

• The student will be able to read specific requested information from a phase diagram. 

• The student will be able to state the primary difference between a generic phase diagram and a phase 
diagram for water. 

Introduction 

A phase diagram is a convenient way of representing the phase of a substance as a function of temperature 
and pressure. Phase diagrams are produced by altering the temperature of a pure substance at constant 
pressure in a closed system. This process is repeated at many different pressures and the resultant phases 
charted. 

Generic Phase Diagram 

The phase diagram in Figure 14.6 is a generic phase diagram that would be produced by many pure 
substances. The differences in the diagram for different substances would be in the specific thermodynamic 
points like melting points, boiling points, and so on and differences in the slopes of the curved lines. The 
general shape of the phase diagram would be very similar for many substances. 





g x 



a 

E 



Solid 

A 


/ Liquid 


T? r 


Gas 




\^ s 



B 

Temp«petuj*« in °C 



Figure 14.6: A generic phase diagram. 

In the pink area in the diagram, the substance would be in the solid state, the purple area represents liquid, 
and in the yellow area, the substance would be gaseous. Following a constant pressure line, such as XY, 
shows the phase of the substance at different temperatures for this pressure. Since line XY crosses from 
solid into liquid at point A, this temperature (B) would be the melting point of the substance. Continuing 
along the line, we see it crosses from liquid to gas and that point corresponds with temperature C. This is 
the boiling point of the substance at pressure X. The line between the pink and purple areas represent the 
various melting points at different pressures and the line separating the purple area from the yellow area 



www.ckl2.org 



410 



represents the boiling point at various pressures. At the melting points, both solid and liquid can exist at 
the same time as the phase changes occurs but on either side of this temperature, the substance must be 
all in one phase. At the boiling points, the substance may exist in both liquid and gas phase at the same 
time but only exactly at the boiling point. There is one point on the diagram where all three phases may 
exist at one time. This point is called the triple point. The pressure at this point is called the triple 
point pressure and the temperature at this point is called the triple point temperature. 

There is also a line separating the pink area from the yellow area. This line represents the phase change 
in which a solid changes directly to a gas without passing through the liquid phase. This phase change is 
known as sublimation. All substances undergo sublimation at the appropriate pressures. We do not see 
sublimation often because the pressures are frequently quite low and we do not encounter substances at low 
pressures in our daily lives. Some of us have seen carbon dioxide, CO2, in the solid form which is called dry 
ice. If you have seen dry ice, you have noticed that the substance goes from the solid phase to the gaseous 
phase at room conditions without passing through a liquid phase. In the phase diagram for dry ice, we 
would see that the triple point is above normal atmospheric pressure and so at room conditions, carbon 
dioxide undergoes sublimation. Water also undergoes sublimation but the pressure where this would occur 
is below 4.58 torr or 0.0060 atm - a pressure we seldom witness. 



Solid 




Liquid 


A i 


TP/ 


Gas 









Temperature in C 

Figure 14.7: Another generic phase diagram. 

Figure 14.7 shows the same generic phase diagram we looked at before. Two points have been added to 
the diagram, labeled A and B. You should note that the substance at point A can be caused to go through 
a phase change from solid to gas (sublimation) in two different ways. The substance could be heated at 
constant pressure or the substance could undergo a lowering of pressure at constant temperature. Both of 
these procedures would cause the solid to undergo sublimation. Point B is a similar circumstance except 
that the substance begins as a liquid. The liquid at point B could be caused to under a phase change to 
gas either by heating it at constant pressure or by lowering the pressure at constant temperature. You 
might also note that the substance at the triple point will become a solid if pressure is increased and will 
become a gas if pressure is decreased. 

Phase Diagram for Water 

The phase diagram for water has one very interesting difference from the generic phase diagram. Please 
note that this diagram is not drawn to scale. If the distance between 1.0 atm and 218 atm was drawn to 
scale, the difference between 1.0 atm and 0.006 atm would be invisible. The diagram is drawn just to show 
specific points of interest. (See Figure 14.8). 

The primary difference in the shape of this diagram and the generic diagram is that the solid-liquid 
equilibrium line (melting point) has a negative slope (tilts backwards). The positive slope of this line in 
the generic diagram indicates that as pressure increases, the melting point increases. That is reasonable 



411 



www.cki2.0rg 



Phase Diagram for Water 
213 



E 

O 

I 

B 

S 

z 

£1- 



1.0 



0.006 




374 



Temperature in C 



Figure 14.8: The phase diagram for water. 



because more pressure on the surface would require a higher temperature to overcome that extra pressure 
and melt the substance. The negative slope of this line in the water diagram indicates that as the pressure 
increases, the melting point of water decreases. The reason this occurs is because the increased pressure 
breaks some of the hydrogen bonds in the water and so LESS temperature is needed to melt ice at higher 
pressures. There is plenty of evidence of this in everyday life if you look for it. We all think of ice as being 
a very slippery substance but the surface of ice is no different from the surface of many other solids. The 
surface of ice is not smoother than the surface of copper or glass or most other solids. The reason that 
we slip on ice is that when you stand on ice or drive your car on ice, the pressure of your weight or the 
weight of your car causes the ice to melt. When you put pressure on ice, the surface between the ice and 
the weight is liquid and that makes it slippery. If you are an ice skater, you are aware that when you look 
closely at the track of the blade of an ice skate on ice, the track is filled with liquid, not solid. 

If you follow the line at 1.0 atm pressure for water, you see that the melting points and boiling points are 
the temperatures noted as the normal melting point and the normal boiling point. The triple point for 
water is at 0.006 atm and 0.0098°C Very expensive equipment is necessary if you wish to see water at 
its triple point. There are commercial processes that make use of the sublimation of water. Foods that 
are referred to as "freeze dried" have the surrounding pressure and temperature reduced to a point below 
the triple point and then they are heated while a vacuum pump removes vapor to keep the pressure below 
the triple point pressure. This causes the water in the food to sublimate (solid to gas) and it is drawn off 
by the pump. The end result is the food minus all the water that was in it. This process is supposed to 
damage the food less than getting rid of the water by heating the food. The idea is that you can just add 
water and the food will be like it was originally. The success or failure of that idea is for you to decide. 

As the temperature of liquid is raised, the amount of pressure that is required to squeeze the substance and 
keep it in liquid form also increases. Liquids will eventually reach a temperature at which no amount of 
pressure will keep it in the liquid form. The substance at that temperature will vaporize regardless of the 
amount of pressure on it. The highest temperature a liquid reaches and can still be squeezed into a liquid 
is called the critical temperature. The pressure that is required at the critical temperature to force the 
liquid to stay in liquid form is called the critical pressure. The critical temperature and pressure for 
water is 374°C and 218 atm. 



www.ckl2.org 



412 



Lesson Summary 



Phase diagrams are graph showing the phase of a substance at various conditions of temperature and 

pressure. 

The phase diagram for water is different from most phase diagrams because unlike most substances, 

the melting point of water decreases as pressure increases. 



Review Questions 

1. Consider the phase diagram below. 




Name the phases that may be present at each lettered point in the diagram. 

Further Reading / Supplementary Links 

• Chemistry, Matter and Its Changes, 4 th Edition, Chapter 12: Intermolecular Attractions and the 
Properties of Liquids and Solids, James E. Brady and Fred Senese, John Wiley &; Sons, Inc., 2004. 
http : //dwb . unl . edu/teacher/nsf/cO 11 inks/www . ualberta . ca/ ! bderksen/f lorin . html 

• http : //www . sciencedaily . com/releases/2008/07/080704153507 . htm 

Vocabulary 

critical pressure The pressure required to liquefy a gas at its critical temperature. 

critical temperature The highest temperature at which it is possible to liquify the substance by in- 
creasing pressure. 



Image Sources 



(1) http://www.openchemistry.co.uk/images/diamond. Creative Commons Attribution 3.0. 

(2) Richard Parsons. Ionic solids are not malleable.. CC-BY-SA. 

413 www.ckl2.ors 



(3) Richard Parsons. The phase diagram for water.. CC-BY-SA. 

(4) Richard Parsons. A NaCl crystal. CC-BY-SA. 

(5) Richard Parsons. The heating curve for water.. CC-BY-SA. 

(6) Richard Parsons. A generic phase diagram.. CC-BY-SA. 

(7) http://en.wikipedia.0rg/wiki/File:AsbestoslUSGOV.jpg. Public Domain- USGov. 

(8) Richard Parsons. Another generic phase diagram.. CC-BY-SA. 



www.ckl2.org 414 



Chapter 15 



The Solution Process 



15.1 What Are Solutions? 

Lesson Objectives 

• Define solutions. 

• Describe the composition of homogeneous solutions. 

• Describe the different types of solutions that are possible within the three states of matter. 

• Identify homogeneous solutions of different types. 

Introduction 

In this chapter we begin our study of solution chemistry. We have previously discussed a number of the 
concepts you will be learning in this chapter. But we will explore these ideas in greater detail in this 
chapter. Let's begin with a discussion of the definition of a solution. We all probably think we know what 
a solution is. We might be holding a can of soda or a cup of tea while reading this lesson and think ... 
hey this is a solution. Well, you are right. But you might not realize that alloys, such as brass, are also 
classified as solutions, or that air is a solution. Why are these classified as solutions? Why wouldn't milk 
be classified as a true solution? To answer these questions, we have to learn some specific properties of 
solutions. Let's begin with the definition of a solution and view some of the different types of solutions. 

Homogeneous Mixtures 

A homogeneous mixture is a solution of the same appearance or composition throughout. Thinking of 
the prefix "homo" meaning "sameness", this definition makes perfectly good sense. Homogeneous solutions 
carry the same properties throughout the solution. Take, for example, vinegar that is used in cooking. 
Vinegar is approximately 5% acetic acid in water. This means that every teaspoon of vinegar that is 
removed from the container, contains 5% acetic acid and 95% water. 

A point should be made here that when a solution is said to have uniform properties throughout, the 
definition is referring to properties at the particle level. Well, what does this mean? Let's consider brass 
as an example. The brass is an alloy made from copper and zinc. To the naked eye this brass coin seems 
like it is just one substance but at a particle level two substances are present (copper and zinc). An 
alloy is a homogeneous solution formed when one solid is dissolved in another. So the brass represents a 

415 www.ckl2.org 



homogeneous mixture. Now, consider a handful of zinc filings and copper pieces. Is this now a homogeneous 
solution? The properties of any scoop of the "mixture" you are holding would not be consistent with any 
other scoop you removed from the mixture. Thus the combination of zinc filings and copper pieces in a 
pile does not represent a homogeneous solution. Another example of a solution is margarine. Margarine is 
a combination of a number of substances at the molecular level but to the naked eye it is a homogeneous 
solution that looks like just one substance. 

Varying Concentrations of Ingredients Produces Different Solu- 
tions 

The point should be made that because solutions have the same composition throughout does not mean 
you cannot vary the composition. If you were to take one cup of water and dissolve 1/4 teaspoon of table 
salt in it, a solution would form. The solution would have the same properties throughout, the particles of 
salt would be so small that they would not be seen and the composition of every milliliter of the solution 
would be the same. But you can vary the composition of this solution to a point. If you were to add another 
1/2 teaspoon of salt to the cup of water, you would make another solution, but this time there would be a 
different composition than the last. You still have a solution where the salt particles are so small that they 
would not be seen and the solution has the same properties throughout, thus it is homogeneous. What 
would happen if you tried to dissolve 1/2 cup of salt in the water. Would the solution stay homogeneous? 
No, it would not. At this point, the solution has passed its limit as to the amount of salt it can dissolve 
and it would no longer be a homogeneous solution. 

So solutions have constant composition but you can vary the composition up to a point. There are limits 
to the amount of substance that can be dissolved into another substance and still remain homogeneous. 

Types of Solutions 

There are three states of matter: solid, liquid, and gas. If we think about solutions and the possibilities of 
combining these states together to form solutions, we have nine possibilities. Look at the Table 15.1. 

Table 15.1: Types of Solutions 

Solid Liquid Gas 

Solid Solid in a Solid Solid in a Liquid Solid in a Gas 

Liquid Liquid in a Solid Liquid in a Liquid Liquid in a Gas 

Gas Gas in a Solid Gas in a Liquid Gas in a Gas 



In the table (15.1), there are really only four that are common types of solutions. These are shown in 
boldface. The others, although still solutions, are less common in everyday lives. For example, a solid 
in a liquid solution can be anything from salt or sugar solution, to seawater. Liquid in liquid solutions 
include the antifreeze/coolant we use for our cars and vinegar. For a gas in a liquid solution, the most 
common example is soda pop: carbon dioxide dissolved in water (with lots of sugar!) Another example is 
the ammonia solution you may use (or have seen used) to clean in the home. Finally, to understand the 
gas in a gas solution, take a deep breath. That's right, air is a solution made up of mostly oxygen gas and 
nitrogen gas. 

A solid in a solid solution is less common but still we see a lot of steel and brass around in our everyday 
world. These are examples of solid - solid solutions. The other types of solutions are less common but do 
exist in the world of solution chemistry. 

www.ckl2.org 416 



Lesson Summary 

• A solution is a mixture that has the same properties throughout. Solutions have the same composition 
throughout but this composition can vary up to a point, or limit. 

• With three states of matter, four types of solution can be classified as the most common as far as 
occurring in the everyday world. These include solid in a liquid, liquid in a liquid, gas in a liquid, 
and gas in a gas. The other types are less common. 

Review Questions 

1. What makes a solution homogeneous? 

2. Which of the following are homogeneous? Explain. 



(a) 


gasoline 




(b) 


chocolate 




(c) 


blood 




(d) 


brass 




Which of the following is 


a solution? 


(a) 


milk 




(b) 


blood 




(c) 


gold 




(d) 


air 




(e) 


sugar 





4. Which of the following is not a true solution? 

(a) vinegar 

(b) sand and water 

(c) hard water, CaCO^aq) 

(d) mercury alloy 

5. Give an example of a homogeneous solution that is made from the following combinations 

(a) a gas in a liquid 

(b) a solid in a solid 

(c) a solid in a liquid 

(d) a gas in a gas 

6. Jack is practicing some household chemistry. He takes 1 tsp of sugar and dissolves it in 250mL of 
water. He sees that the solution remains clear so continues his experiment by adding a second tsp of 
sugar. Stirring the solution makes this solution turn clear. After a few more attempts, Jack sees the 
solution turn murky then sugar crystals sinking to the bottom. What is Jack demonstrating? 

Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

solution A homogenous mixture; composition can vary; but composition is the same throughout once 
the solution is made. 

417 www.ckl2.org 



15.2 Why Solutions Occur 

Lesson Objectives 

• Describe why solutions occur; the "like dissolves like" generalization. 

• Determine if solutions will occur by studying the molecular structure. 

• State the importance of water as the "universal solvent." 

Introduction 

We have learned that solutions can be formed in a variety of combinations using solids, liquids and gases. 
We also know that solutions have constant composition and we can also vary this composition up to a 
point to maintain the homogeneous nature of the solution. But how exactly do solutions form? Why is it 
that when you mix oil and water together a solution does not form even though it is a liquid in a liquid 
and yet vinegar and water will? Why could we dissolve table salt in water but not in vegetable oil? The 
reasons why solutions form will be explored in this lesson, along with a discussion of why water is used 
most frequently to dissolve substances of various types. 

Here is some new vocabulary you will meet in this chapter. 

Solution: a homogeneous mixture of substances 

Solute: the substance dissolved in a solution, usually determined by being the smaller quantity 

Solvent: the substance the solute is dissolved in, usually determined by being the larger quantity 

Similar Structures Allow Solutions to Occur 

Over the course of your study in chemistry you have learned the terms polar and non-polar. Recall that 
in chemistry, a polar molecule is one that has a positive end and a negative end while nonpolar molecules 
have charges that are evenly distributed throughout the molecule. In fact, during the study of Valence 
Shell Electron Pair Repulsion Theory (VSEPR), you learned that the chemical structures themselves have 
built in molecular polarity. 

In solution chemistry, we can predict when solutions will form and others won't using a little saying ... "like 
dissolves like". The "like dissolves like' saying helps us to predict solubility based on the two parts of a 
solution having similar intermolecular forces. For example, suppose you are dissolving methanol in water. 
Both methanol and water are polar molecules and form a solution because they both have permanent 
dipoles (positive and negative parts of the molecules) that allow the molecules of each of the substances to 
be attracted to the other. When this occurs, a solution is made. 

A way to understand this is to think about why Velcro is used to hold two different pieces of fabric together. 
The two sides of Velcro allow the pieces of fabric to be fastened together because the Velcro has similar 
structure that "attract" each other. However, one side of Velcro would not stay together with a piece of 
silk since the silk doesn't have any part of its structure with which the Velcro can connect. 

Let's look at the individual structure of the water and methanol molecules. Notice in the representation 
of the individual molecules of methanol and water how the methanol has a permanent dipole due to the 
C — O — H bonds and is a polar molecule. In the representation of the water molecule, you can see that 
there are also permanent dipoles making it a polar molecule. The intermolecular forces for both of these 
molecules are dipole-dipole attractions. Since these molecules are both polar, they will form a solution 
when mixed together. We say they are miscible, which means these two liquids will make a solution. (See 
Figure 15.1). 

www.ckl2.org 418 




4H^ 




sodium chloride crystal 



hydrated sodium ion 



hydrated chloride ion 




The more electronegative oxygen atom pulls the shared 
electrons away from the hydrogen atoms causing an 
unequal distribution of electrons over the water 
molecule. The hydrogen end of the water molecule will 
be slightly negative and the oxygen end will be slightly 
positive. A molecule with this permanent uneven 
distribution of electrons is said to be polar. 

When a polar or ionic compound is introduced into 
water : ionic charges on the ions or the poles on a polar 
molecule are attracted to the poles on the water 
molecule and the substance dissolves. 



Figure 15.1: The hydration of ions in a polar solvent. 

The same is true for the case of a non-polar substance such as carbon tetrachloride being dissolved in 
another nonpolar substance such as pentane. London-dispersion forces are the intermolecular bonds that 
hold the carbon tetrachloride together as a liquid. London-dispersion forces are also the forces that allow 
pentane to be a liquid at room temperature. Since both of these substances have the same intermolecular 
forces, when they are mixed together, a solution will be formed. 

Unlike the polar molecules, the non-polar molecules have, at any given time, no permanent dipole. If we 
were to add table salt, NaCl, to either carbon tetrachloride or pentane, we would find that the salt would 
not dissolve. The reason for this is again explained with the structures of the substances. Since carbon 
tetrachloride (or pentane) has no permanent dipoles in its molecules, there would be place for the charges 
particles in a crystal of NaCl to be attracted. 

In a polar solvent, the molecules of solvent are attracted to each other by the partial charges on the ends 
of the molecules. When a polar solute is added, the positive polar ends of the solute molecules attract the 
negative polar ends of the solvent molecules and vice versa. This attraction allows the two different types 
of molecules to form a solution. If a non-polar solute was added to a polar solvent, the non-polar solute 
particles cannot attract the solvent molecules away from each other - so a solution does not form. 

Polar solvents will dissolve polar and ionic solutes because of the attraction of the opposite charges on the 
solvent and solute particles. Non-polar solvents will only dissolve non-polar solutes because they cannot 
attract the dipoles or the ions. 



Water: The Universal Solvent 

Think of the title of this section, water: the universal solvent. The term solvent is used to represent the 
medium that is used to produce the solution. The term universal is used to describe the fact that water, 
along with many of its other unique aspects, can dissolve many types and kinds of substances. For instance, 
table salt, NaCl, is an ionic compound but easily makes a solution with water. This is true for many ionic 
compounds. And from your own experience you know that table sugar, a polar covalent compound, also 
dissolves in water. And this is also true for other polar compounds such as vinegar and corn syrup. 



419 



www.ckl2.org 



Even some nonpolar substances dissolve is water but only to a limited degree. Have you ever wondered 
why fish are able to breathe? Oxygen gas, a nonpolar molecule, does dissolve in water and it is this oxygen 
that the fish take in through their gills. Or, one more example of a nonpolar compound that dissolves in 
water is the reason we can enjoy carbonated sodas. Pepsi-cola and all the other sodas have carbon dioxide 
gas, CO2, a nonpolar compound, dissolved in a sugar-water solution. In this case, to keep as much gas in 
solution as possible, the sodas are kept under pressure. But that's another part of the story! 

Lesson Summary 

• Whether or not solutions are formed depends on the similarity of polarity or the "like dissolves like" 
rule. Polar molecules dissolve in polar solvents, non-polar molecules dissolve in non-polar solvents, 
and ionic molecules in polar solvents. Polarity is determined from molecular geometry. 

• Water is considered as the universal solvent since it can dissolve both ionic and polar solutes, as well 
as some non-polar solutes (in very limited amounts). 

Combinations to form solutions with polar, ionic, and non-polar substances in Table 15.2. 

Table 15.2: 

Combination Solution formed 

Polar substance in a polar substance yes 

Non-polar substance in a non-polar substance yes 

Polar substance in a non-polar substance no 

Ionic substance in a polar substance (i.e. water) yes 

Ionic substance in a non-polar substance no 



Review Questions 

1. What is the "like dissolves like" generalization and provide an example to illustrate your answer. 

2. Why will LiCl not dissolve in CC/4? 

3. Will acetic acid dissolve in water? Why? 

4. What is the difference between intermolecular and intramolecular bonds? 

5. In which compound will benzene (CqHq) dissolve? 

(a) Carbon tetrachloride 

(b) water 

(c) vinegar 

(d) none of the above 

6. In which compound will sodium chloride dissolve? 

(a) Carbon tetrachloride 

(b) methanol 

(c) vinegar 

(d) none of the above 

7. In which compound will ammonium phosphate dissolve? 

(a) Carbon tetrachloride 

(b) water 

(c) methanol 

www.ckl2.org 420 



(d) None of the above 

8. Thomas is making a salad dressing for supper using balsamic vinegar and oil. He shakes and shakes 
the mixture but cannot seem to get the two to dissolve. Explain to Thomas why they will not 
dissolve. 

Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

intermolecular bonds Forces of attraction between molecules. 
intramolecular bonds Forces of attraction between atoms in a molecule. 
universal solvent A solvent able to dissolve practically anything (water). 



15.3 Solution Terminology 

Lesson Objectives 



• Define solute, solvent, soluble, insoluble, miscible, immiscible, saturated, unsaturated, concentrated, 
and dilute. 

Introduction 

Like any discipline that you would study or learn in our world, there are terms that are a part of the 
normal day- to- day conversation that takes place. The same is true when people get together to talk 
about solutions. You cannot talk about skateboarding without knowing about grinds and slides; you can't 
talk about bowling without knowing about spares and strikes; and, you can't talk about solutions without 
knowing terms such as solute, dilute, and saturated, just to name a few. In this section, you will learn the 
terms of solution chemistry. 

Solvent and Solute 

The solvent and solute are the two basic parts of a solution. The term solvent is the substance present 
in the greatest amount. The solute, then, is the substance present in the least amount. Let's think for a 
minute that you are making a cup of hot chocolate. You take a teaspoon of chocolate and dissolve it in 
one cup of hot water. Since the chocolate is in the lesser amount it is said to be the solute; and the water 
is the solvent since it is in the greater amount. 

Sample question: Name the solute and solvent in each of the following mixtures 

(a) Kool-aid 

(b) iced tea 

(c) soft drinks 

421 www.ckl2.org 



Solutions: 

(a) solute = Kool-aid crystals (and sugar); solvent = water 

(b) solute = ice tea mix (and sugar); solvent = water 

(c) solute = carbonic acid (and sugar or sweetener, chemicals, etc.); solvent = water 

Soluble and Insoluble 

Did you ever try to make a solution and yet it did not matter how much shaking and stirring that you 
did, the solid just kept falling to the bottom of the container. Try it for yourself. Take some sand and try 
to dissolve it in a cup of water. What happens? The sand will not dissolve. It is insoluble. Now if you 
were to take a teaspoon of table salt or sugar and do the same experiment, what a different result. There's 
no problem dissolving these substances. Salt and sugar are both soluble in water. When a substance is 
soluble, it means that the substance has the ability to dissolve in another substance. And being insoluble 
means that the substance does not dissolve. 

Miscible and Immiscible 

When referring to liquid solutes in liquid solvents, we can use the terms miscible and immiscible. Liquids 
are said to be miscible when they can dissolve in each other. Therefore if they are immiscible, they are 
insoluble. When making a cake using a cake mix, you often use 1+ cups of water and 1/3 cup of oil and 
then you add this to the mix. If you first mix this together you would see that the mixture is an immiscible 
solution. The oil does not dissolve in the water. Since cooking oil is non-polar and the water is polar, they 
have different types of intermolecular bonds and they will not make a solution. When making biscuits, on 
the other hand, the recipe may call for you to add a little vinegar to the water before adding the liquid 
to the dry ingredients. The vinegar will be miscible in water because both vinegar and water are polar 
compounds and therefore have the same type of intermolecular bonds and can make a solution. 

Saturated and Unsaturated 

A saturated solution is one in which a given amount of solvent has dissolved the absolute maximum solute 
at that temperature. Let's go back to our table salt and water example from before: now try to dissolve 
2 teaspoons of table salt in one cup of water. This is probably the maximum amount of salt that could 
dissolve. Try now to dissolve 3 teaspoons, some of the table salt would probably sit on the bottom of the 
glass; but at 2 teaspoons, all of the salt is in solution. The solution becomes saturated. If more solute 
is added to a saturated solution, the excess solute remains undissolved and simply sits on the bottom of 
the cup. If only one (1) teaspoon was placed in the glass, the solution would be said to be unsaturated. 
This is because the solvent is holding less than the maximum amount of solute at that temperature. An 
unsaturated solution is one that contains less than the maximum amount of solute that is possible in a 
given amount of solvent. 

Concentrated and Dilute 

Solutions can also be said to be dilute or concentrated. A dilute solution is a concentrated solution that has 
been, in essence, watered down. Think of the frozen juice containers you buy in the grocery store. What 
you have to do is take the frozen juice from inside these containers and usually empty 3 or 4 times the 
container size full of water to mix with the juice concentrate and make your container of juice. Therefore, 
you are diluting the concentrated juice. When we talk about solute and solvent, the concentrated solution 

www.ckl2.org 422 



has a lot of solute verses the dilute solution that would have a smaller amount of solute. A concentrated 
solution is one in which there is a large amount of solute in a given amount of solvent. A dilute solution 
is one in which there is a small amount of solute in a given amount of solvent. 

Lesson Summary 

• Generally speaking, in a solution, a solute is present in the least amount (less than 50% of the 
solution) whereas the solvent is present in the greater amount (more than 50% of the solution). 

• When a substance can dissolve in another it is said to be soluble; when it cannot, it is said to be 
insoluble. For two liquids, when they are soluble in each other the liquids are said to be miscible; 
when they are insoluble the liquids are considered immiscible. 

• A saturated solution holds the maximum amount of solid at a specific temperature. An unsaturated 
solution does not have the maximum amount of solute dissolved at that temperature in a given 
amount of solvent. A concentrated solution has a large amount of solute in a given amount of 
solvent. A dilute solution has a lesser amount of solute. 

Review Questions 

1. Distinguish between soluble, insoluble and miscible, immiscible. Use an example in your answer. 

2. How can a solution that is concentrated be made more dilute and a dilute be made more concentrated? 

3. Vinegar and water will mix together. Therefore two liquids are said to be: 

(a) saturated 

(b) miscible 

(c) unsaturated 

(d) immiscible 

4. A solution is analyzed and found to contain 90 g of solute in 100 mL of solution. What can be 
concluded about this solution? 

(a) The solution is concentrated. 

(b) The concentration of the solution is 90 g/100 mL of water. 

(c) The solution is saturated. 

(d) The solution is holding the maximum amount of solute. 

5. A solute is defined as: 

(a) The substance in a solution present in the least amount. 

(b) The substance in a solution that represents less than 50% of the solution. 

(c) The substance that is dissolved in the solvent. 

(d) All of the above. 

(e) None of the above. 

6. Match the following words with the examples that describe them. 

(a) solute - Adding only one can of water to a frozen concentrated juice mix will form this type of 
solution. 

(b) solvent - Adding eight cans of water to a frozen concentrated juice mix will form this type of 
solution. 

(c) soluble - Alcohol and water will have this property. 

(d) insoluble - Gasoline and water will have this property. 

(e) miscible - The water of a NaOH{aq) solution. 

(f) immiscible - When salt is added to water it is said to have this property. 

(g) saturated - The copper(II) sulfate crystals in a solution of CuS 0±(aq) 

423 www.ckl2.org 



(h) unsaturated - The maximum amount of silver nitrate that can dissolve in 100 mL of water is 

220 g. What term is given to this solution? 
(i) concentrated - If 220 g of AgNO% can dissolve in 100 g of water and only 50 g are added, what 

type solution is formed? 
(j) dilute - Adding calcium hydroxide to water forms a milky white precipitate. What term is given 

to calcium hydroxide? 

7. Nisi is given two bottles of copper (II) sulfate solutions in her senior high chemistry lab. She is told 
that one bottle contains a saturated solution and the other one contains an unsaturated solution. 
What can Nisi do to identify the two solutions? 

8. Can you have a solution that is saturated and dilute at the same time? Explain. 

Vocabulary 

solute The substance in a solution present in the least amount. 

solvent The substance in a solution present in the greatest amount. 

soluble The ability to dissolve in solution. 

insoluble The inability to dissolve in solution. 

miscible Two liquids having the ability to be soluble in each other. 

immiscible Two liquids not having the ability to be soluble in each other. 

saturated A solution holding the maximum amount of solution in a given amount of solvent. 

unsaturated A solution holding less than the maximum amount of solution in a given amount of solvent. 

concentrated A solution where there is a large amount of solute in a given amount of solvent. 

dilute A solution where there is a small amount of solute in a given amount of solvent. 

15.4 Measuring Concentration 

Lesson Objectives 

• Define molarity, mass percent, ppm, and molality. 

• Calculate molarity, mass percent, ppm, and molality. 

• Explain the importance of quantitative measurement in concentration. 

Introduction 

Although qualitative observations are necessary and have their place in every part of science, including 
chemistry, we have seen throughout our study of science that there is a definite need for quantitative 
measurements in science. This is particularly true in solution chemistry. We might read in the headlines 

www.ckl2.org 424 



that the amount of mercury found in the fish is up by 0.5ppm and say to ourselves, what does that mean? 
Is it important? We read labels in the grocery store that are in weight percent. What does this mean? So 
being able to deal with the quantitative side of solutions helps us to move toward a deeper understanding 
of solutions, one that involves not only a numerical analysis but a critical analysis as well. Let's explore 
some of the different quantitative applications of solution chemistry. 

Molarity 

Of all the quantitative measures of concentration, molarity is the one used most frequently by chemists. 
Molarity is defined as the moles of solute per liter of solution. The symbol given for molarity is M. You 
will see molarity units of both moles/liter and M. Chemists also used square brackets [ ] to indicate a 
reference to the molarity of a substance. For example, the expression [Ag + ] refers to the molarity of the 
silver ion. 

It should be noted that when making solutions to a certain molarity, the amount of solvent to be used 
cannot be measured. The amount of solvent used will be whatever is necessary to bring the total solution 
to the required volume. This amount of solvent cannot be calculated beforehand either by volume or by 
mass because it is not known what volume will be required to reach the total volume of solution. Solution 
concentrations expressed in molarity are the easiest to calculate with but the most difficult to make in the 
lab. 

moles of solute 
molarity = mols/L 



liters of solution 



Sample question 1: What is the concentration, in mol/L, where 2.34 mol of NaCl has been dissolved in 
500 mL of H 2 0. 

Solution: 

[NaCl] = 2M m ° 1S = 4.68M 
L J 0.500 liters 

The concentration of the NaCl solution is 4.68 mol/L 

Sample question 2: What would be the mass of 500 mL of a 1.25 mol/L potassium sulfate solution? 

Solution: 

mol 
M = so mols M x L 

mols = (1.25 mol/L(0.500L) = 0.625 mol 

mols = so mass = (mols) (molar mass) 

molar mass 

mass = (0.625 mol) (174.3 g/mol) = 109 g 
Therefore the mass of K2SO4 that dissolves in 500 mL of H2O to make this solution is 109 g. 

Mass Percent 

Mass percent, is the number of grams of the solute in the number of grams of solution. Mass percent is a 
term frequently used when referring to solid solutions. It has the formula: 

mass of solute 

percent by mass = X 100 

solute mass + solvent mass 



425 www.ckl2.org 



or 

mass solute 

percent by mass = X 100 

mass solution 

Sample question: An alloy is prepared by adding 15 g of zinc to 65 g of copper. What is the mass percent 
of zinc? 

Solution: 

mass of solute 

percent by mass = X 100 

solute mass + solvent mass 

mass Zn 15 g 15 g _, 

percent by mass = — x 100 = x 100 = — - x 100 = 19% 

mass Zn + mass Cu 15 g + 65 g 80 g 

Parts Per Million 

Parts per million is another unit for concentration. Obscure in the sense that it is used less frequently 
than the others that will be dealt with in this unit. This does not mean that it is any less important 
than the others, it just means that for the normal "day to day" conversations that folks have around the 
solution chemistry lab, parts per million might not come up all that often. Parts per million denotes that 
there is 1 milligram of solute for every kilogram of solvent. Usually you hear about parts per million when 
governments are talking to us about drinking water or poisons in fish and other food products. It is used 
most frequently when dealing with environmental issues when you pick up the newspaper or common news 
magazines around the house. To calculate parts per million, the following formula is used. 

mass of solute R 

ppm = t — rr^~ x 10 

mass oi solution 

Sample question: Mercury levels in fish have often been at the forefront of the news for people who love to 
eat fresh fish. Salmon, for instance, contains O.Olppm compared to shark which contains 0.99 ppm. In the 
United States, canned tuna is the most popular selling fish and has a mercury level of 0.12 ppm according 
to the FDA statistics. If one were to consume 1.00kg of canned tuna over a certain time period, how much 
mercury would be consumed? 



Solution 



mass of solute p 

ppm = t — rr^~ x 10 

mass oi solution 



so 



mass of solute 



(mass of solution) (ppm) 

1 x 10 6 
(1000. g)(0.12) 10 _ n _ 4 



mass of solute = 7 . = 1.2 x 10 

1 x 10 6 

Molality 

Molality is one further way to measure concentration of a solution. It is calculated by dividing the moles 
of solute by the kilograms of solvent. Molality has the symbol, m. 

. ,. , . moles solute 

molality (m) = 

kg of solvent 

www.ckl2.org 426 



An interesting sidebar: molality is used for specific topics in solution chemistry as molarity as a method for 
measuring concentration. So why learn both? Well, oddly enough it all boils down to temperature ... pardon 
the pun! Molarity, if you recall, is moles of solute per volume of solution and the volume is temperature 
dependent. As the temperature rises, the molarity of the solution will actually decrease slightly because the 
volume will increase slightly. Molality does not involve volume and mass is not temperature dependent. 
Thus, there is a slight advantage to using molality over molarity when temperatures move away from 
standard conditions. Yet, still, the choice of majority falls to using molarity. 

Sample question: Calculate the molality of a solution of hydrochloric acid where 12. 5g of hydrochloric acid 
has been dissolved in 115 g of water. 

Solution: 

12 5 e: 

mol HC1 = —^ — = 0.343 mol 

36.46 g/mol 

TT ^, moles solute 0.343 mol 

molality HC1 = = — = 2.98 m 

kg solvent 0.115 kg 

Lesson Summary 

• Molarity is the moles of solute per liter of solution. Molarity normally uses the symbol M. Mass 
percent is the number of grams of the solute in the number of grams of solution, of course multiplied 
by 100. 

• Parts per million means that there is 1 milligram of solute for every kilogram of solvent. Therefore 
it is the mass of solute per mass of solution multiplied by 1 million. 

• Molality is calculated by dividing the moles of solute by the kilograms of solvent. It is less common 
than molarity but more accurate because of its lack of dependence on temperature. 

Review Questions 

1. Calculate the mass percent of silver when a silver /nickel solution is made with 34.5 g of silver and 
72.3 g of nickel. 

2. What would be the ppm of silver for the data presented in question 1? 

3. Why is it a good idea to learn mass percent when molarity and molality are the most commonly used 
concentration measures? 

4. Most times when news reports indicate the amount of lead or mercury found in foods, they use the 
concentration measures of ppb (parts per billion) or ppm (parts per million). Why use these over 
the others we have learned? 

5. What is the molarity of a solution prepared by dissolving 2.5 g of LiNO% in sufficient water to make 
60 mL of solution? 

(a) 0.036 mol/L 

(b) 0.041 mol/L 

(c) 0.60 mol/L 

(d) 0.060 mol/L 

6. A solution is known to have a concentration of 325 ppm. What is the mass of the solute dissolved in 
1.50 kg of solvent? 

(a) 0.32 mg 

(b) 0.49 mg 

(c) 325 mg 

(d) 488 mg 

427 www.ckl2.org 



7. Calculate the molality of a solution of copper (II) sulfate where 11.25 g of the crystals has been 
dissolved in 325 g of water. 

(a) 0.0346 m 

(b) 0.0705 m 

(c) 0.216 m 

(d) None of the above 

8. What is the mass of magnesium chloride present in a 250 g solution found to be 21.4% MgCl<p. 

(a) 21.4 g 

(b) 53.5 g 

(c) 196.5 g 

(d) 250 g 

9. What is the concentration of each of the following solutions in mol/L. 

(a) 3.50 g of potassium chromate dissolved in 100 mL of water 

(b) 50.0 g of magnesium nitrate dissolved in 250 mL of water. 

10. Find the mass of aluminum nitrate required to produce 750 g of a 1.5 molal solution. 

11. The Dead Sea contains approximately 332 grams of salt per kilogram of seawater. Assume this salt 
is all NaCl. Given that the density of the Dead Sea water is approximately 1.20 g/mL, calculate: 

(a) the mass percent of NaCl. 

(b) the mole fraction of NaCl. 

(c) the molarity of NaCl. 

Further Reading / Supplemental Links 

Website with lessons, worksheets, and quizzes on various high school chemistry topics. 

• Lesson 6-4 is on Molarity. 

• http : //www . f ordhamprep . org/gcurran/sho/sho/lessons/lesson31 . htm 

Vocabulary 

molarity A concentration unit measuring the moles of solute per liter of solution. 

mass percent A concentration unit measuring the mass of solute per mass of solution. This unit is 
presented as a percent. 

weight percent Another name for mass percent. 

parts per million (ppm) A concentration unit measuring the mass of solute per mass of solution multi- 
plied by 1 million. 

molality A concentration unit measuring the moles of solute per kilograms of solutions. 
www.ckl2.org 428 



15.5 Solubility Graphs 

Lesson Objectives 

• Define solubility. 

• Read and report data from solubility graphs. 

• Read and report saturation points from a solubility graph. 

Introduction 

In an earlier chapter we have discussed solubility as it applied to data analysis. Solubility graphs are 
an excellent way of organizing and displaying data for interpretation. In this lesson, we will explore this 
concept in learning how to read and analyze a solubility graph in order to extract the relevant data. In 
the everyday world, solutions play a key role in our lives from the foods we eat (through proper mixing), 
the solutions we prepare to clean homes, and also are essential in the laboratory. 

Reading Solubility From a Graph 

Solubility, as we already know, is the amount of solute that will dissolve in a given amount of solvent 
at a particular temperature. The latter part of this statement is significant since, for many solutes, the 
solubility will increase as the temperature is increased. There are exceptions, of course, just as there are 
exceptions to every rule. Sodium chloride (table salt), for example, will dissolve to the extent of about 36 g 
in 100 g of water at 25°C and there is little change as the temperature increases. The solubility of cesium 
sulfate actually decreases as temperature increases. But a vast majority of ionic solids that are solutes do 
increase their solubility with temperature. 

Think of the solutes you would be dissolving in your own environment. If you make a cup of instant coffee 
you most likely boil the water first to dissolve more of the coffee crystals. If you are making hot chocolate, 
you would also use boiling water; hot tap water would not do the trick quite so well. Making a bowl of 
"Cup of Noodles" simply does not taste quite as good when you do not have the noodles heated in the 
boiling water because the noodles cannot dissolve all of the spices to the flavor that is expected. Think 
also of a can of cola. This is a solution of a gas, CO2, in a liquid. When you have a can of soda at room 
temperature there is more gas above the surface of the liquid in the can than if you have a can of soda 
cooled to an icy cold temperature. This is because the solubility of the gas in solution decreases as the 
temperature is increased and the gas moves up to the air space in the can above the liquid. Think of this 
before you open up the next warm pop. 

To display the different solubilities at different temperatures, a solubility graph is drawn to show the data 
in a more coherent manner. Having a solubility graph allows us to read the data about a particular solute 
or to compare solutes at a particular temperature quickly and easily. Let's look at a typical solubility 
graph and see how it works. 

(See Figure 15.2). 

What kind of information does this graph tell us? You can see that three of the four solids increase 
solubility with increasing temperature, NaCl only slightly, and KNO3 solid increases substantially with 
increasing temperature. In addition to general trends in the solubility of a substance, you can also get 
detailed facts from a solubility graph. For example, we can see that at 30°C, 95 g of sodium nitrate, 
NaNOs, will dissolve but at 60°C, 120 g will dissolve in 100 g of H2O. At these same two temperatures, 
only 50 g of NCI2SO4 and 113 g, of potassium nitrate, KNO3, will dissolve in 100 g of H2O. 

429 www.cki2.0rg 




Figure 15.2: A solubility graph for some common compounds. 

Sample question 1: Answer the following questions using Figure 15.2. 

(a) How much sodium nitrate will dissolve at 30°C? 

(b) What solid is more soluble at 60°C? 

(c) What solid is least soluble at 40° C? 

(d) At what temperature will 60g of sodium sulfate dissolve in 100 g of water? 
Solution: 

(a) Looking at the solubility graph (see Figure ??), you draw a line up (vertically) from 30°C until it 
hits the NaNOz line. Following this, carry the line over (horizontally) to find the amount of NaNO^ that 
dissolves. 

Therefore approximately 95 g of NaNO^ will dissolve in 100 g of water at 30° C. 

(b) The highest line at 60° C is the green line (NaNO^), therefore it is the most soluble at 60° C. 

(c) The lowest line at 40°C is the purple line (NaCl), therefore NaCl is the least soluble at 40°C 

(d) Looking at the solubility graph (see Figure 15.4), you draw a line over (horizontally) from 60 g until 
it hits the Na2SO^ line. Following this, carry the line down (vertically) to find the temperature at which 
60 g of Nci2SOa will dissolve. 

Therefore 60 g of Na2SO^ will dissolve in 100 g of water at 50°C 

Reading Saturated Solutions From a Graph 

Look at the solubility graph that shows more common ionic compounds. The lines on the solubility curves 
represent the amounts that dissolve in the given amount of solvent at a specific temperature. Look at 
the line for NH3. According to the graph that NH3 is the only substance of this group that decreases 
in solubility as the temperature is increased. We can also see that the most soluble substance at room 
temperature (25°C) is NH3 because it is the line highest up on the graph at 25°C The highest point on 



www.ckl2.org 



430 




Figure 15.3: Reading a solubility graph. 




Figure 15.4: Reading a solubility graph. 



431 



www.ckl2.org 



the solubility curve is at approximately y = 92. We can say then that the most soluble substance at 0°C 
is ammonia with a solubility of approximately 92 g in 100 g of water. 

(See Figure 15.5). 



Solubility «i 

i 1 islu t« 1C.-C! 




Figure 15.5: A solubility graph for common ions and ammonia. 

All of this information can be obtained from reading the solubility graph. What other information can you 
obtain from a solubility graph? You could do a number of different types of calculations. For example, 
what if you were doing an experiment in the lab (at room temperature) and needed a saturated solution 
of potassium chloride dissolved in 35 g of water. How much KCl would you need? 

At 25°C (room temperature), approximately 35 g of KCl will dissolve in 100. g of water. For 35 g of water: 

Proportion 

x g KCl 35 g KCl 



35 g H 2 100. g H 2 
x = 12g KCl 

Another type of problem that can be answered using a solubility graph is to determine if you have saturated 
solutions or not. For example, you have a solution of potassium chlorate that you know is 76 g dissolved in 
250 g of water. You want to know if this solution is saturated or unsaturated when your solution is being 
heated at 80° C. 

Looking at the solubility graph, at 80°C, 44 g of KC10 3 will dissolve in 100 g of H 2 0. Therefore we can 
use the same type of equation as used previously to determine how much would dissolve in 250 g of H 2 0. 

Proportion 

x g KC10 3 44 g KC10 3 



250. g H 2 

x 



100. g H 2 
HO.g KC10 3 



Since it is possible to dissolve 110 g of KC10 3 in 250 g of H 2 and our solution only has 76 g dissolved in 
250 g of H 2 0, the solution is unsaturated. 



www.ckl2.org 



432 



Lesson Summary 

• Solubility is the amount of solute that will dissolve in a given amount of solvent at a specific temper- 
ature. A solubility graph is drawn to display the solubility at different temperatures. It is the mass 
of the solute/100 g of H^O versus temperature in °C. 

• From reading a solubility graph, one can determine the mass of solute that can dissolve at specific 
temperatures and also compare solubility's of different substances at specific temperatures. 

Review Questions 

1. Using the graph below, determine: 

(a) How much ammonia will dissolve at 30° C? 

(b) What solid is more soluble at 50° C? 

(c) What solid is least soluble at 60°C? 

(d) At what temperature will 50 g of ammonium chloride dissolve in 100 g of water? 



Soluburcr 

<;'' ■■■■'■' I !'"■: 



NHjCI 



^-f-~"""'"T KCICtj 


L^-^T~ -^-s^"* KCI 


J— — "j""" ^f 




\j^ 









T*mp4rMUr*^C| 



2. Why are solubility graphs useful? 

3. Define solubility and solubility graph. 

4. How many grams of NaCl are in 450 g of water at 30°C if the solubility is 39.8 g per 100 g of water? 

(a) 8.84 g 

(b) 39.8 g 

(c) 100 g 

(d) 179 g 

5. How many moles of ammonium chloride are in 225 g of water at 40°C if the solubility is 45.8 g per 
100 g of water? 

(a) 0.86 mol 

(b) 1.92 mol 

(c) 20.3 mol 

(d) 103 mol 

6. How many moles of potassium chloride are in 500 g of water at 80°C if the solubility is 51.3 g per 
100 g of water? 

(a) 0.140 mol 

(b) 0.688 mol 

(c) 3.44 mol 



433 



www.ckl2.org 



(d) 10.3 mol 

7. Plot the following data on a solubility graph and then answer the questions below. 

(a) Which substance is the most soluble at 50°C? 

(b) Which substance is the least soluble at 90°C? 

(c) What is the solubility of NH 4 C10 4 at 30°C? 

(d) How many grams of NH 4 C10 4 would dissolve in 250 mL at 30° C? 

(e) At what temperature will 20 g potassium sulfate dissolve in 100 g of water? 

Table 15.3: 

Temp (°C) g NH 4 Br/100 g H 2 g NH 4 ClO 4 /100 g H 2 g NaClO 3 /100 g H 2 

60.0 13.0 80.0 

20 75.5 23.5 98.0 

40 92.0 36.8 118.0 

60 107.8 51.5 143.0 

80 126.0 67.9 172.0 

100 146.0 87.0 207.0 

8. Plot the following data on a solubility graph and then answer the questions below. 

(a) Which substance is the most soluble at 50°C? 

(b) Which substance is the least soluble at 90°C? 

(c) What is the solubility of CuS0 4 at 30°C? 

(d) At what temperature will 20 g potassium sulfate dissolve in 100 g of water? 

Table 15.4: 

Temp (°C) g NaCl/ 100 g H 2 g K 2 SO 4 /l00 g H 2 g CuSO 4 /100 g H 2 

35.7 7.4 14.3 

20 36.0 11.1 20.7 

40 36.5 14.8 28.7 

60 37.3 18.2 40.0 

80 38.1 21.4 56.0 

100 39.2 24.1 80.0 



Vocabulary 



solubility The amount of solute that will dissolve in a given amount of solvent at a particular tempera- 
ture. 



solubility graph A solubility graph is drawn to display the solubility at different temperatures. It is 
the mass of the solute/100 g of H 2 versus temperature in °C. 



www.ckl2.org 434 



15.6 Factors Affecting Solubility 

Lesson Objectives 



Describe the factors that affect solid solubility. 
Describe the factors that affect gas solubility. 
Describe how pressure can affect solubility. 



Introduction 

Solubility, as we have learned, is the maximum amount of a substance that will dissolve in a given amount 
of solvent at a specific temperature. There are two direct factors that affect solubility, temperature and 
pressure. The effects of temperature on solubility can be found for both solids and gases but pressure 
effects are only related to the solubility of gases. Surface area, while not affecting how much of a solute 
will be dissolved, is a factor in how quickly or slowly the saturation point will be reached. In this section, 
we will explore all three of these factors and how they affect the solubility of solids and gases. 



The Effect of Temperature on Solubility 

As we learned in the solubility graphs from the previous section, temperature has a direct effect on solubility. 
For the majority of ionic solids, increasing the temperature increases how quickly that solution can be made. 
As the temperature increases, the particles of the solid move faster which increases the chances that they 
will interact with more of the solvent particles. This results in increasing the rate at which a solution 
occurs. 

However, temperature can also increase the amount of solute that can be dissolved in a solvent. Generally 
speaking, as the temperature is increased, more solute particles will be dissolved. Do take note that this 
is a generalization but one with which you might already be familiar. For instance, when you add table 
sugar to water, a solution is quite easily made. When you heat that solution and keep adding sugar, you 
find that large amounts of sugar can be added as the temperature keeps rising. This is how candy is made. 
The reason this occurs is that as the temperature increases, the intermolecular bonds can be more easily 
broken which allows more of the solute particles to be attracted to the solvent particles. 

There are other examples, though, when increasing the temperature has very little effect on how much 
solute can be dissolved. Table salt is a good example: you can dissolve just about the same amount of 
table salt in ice water as you can in boiling water. 

Chemistry often has exceptions to the rules. For all gases, as the temperature increases, the solubility 
decreases. Using the kinetic molecular theory to explain this phenomenon, as the temperature increases, 
the gas molecules move faster and are then able to escape from the liquid. Their solubility then decreases. 

Look at the graph below, ammonia gas, NH3, shows a sharp decline in solubility as the temperature 
increases whereas all of the ionic solids show an increase in solubility as the temperature increases. 



Solubility graph for ionic solids and NH3. 

435 www.cki2.0rg 





-j 


k 


















































m 














NHjCI 






N. ^_^ x -~-*~"^ 


KClOb 




(S soliri*-l«g 




KCI 








































» 






~-~~\ 






"~— -|-^_ 






! ^^ 








K 


» 40 K « tM 
TfeHMHtflirVfCi 




J 



























Remember from earlier chapters, that endothermic changes absorb energy and exothermic changes release 
energy. When gases dissolve in water, it is usually with an exothermic change. Notice that in Figure 1, the 
NH$(g) solubility decreases with temperature and as we look at the equation for the dissolution of NH^(g), 
we see that it is exothermic. 

NH 3(g) -* NH 3(aq) + energy 

A graph for the solubility of oxygen gas, O2, would be very similar to the one for NH^(g); in other 
words, oxygen gas would decrease in solubility as the temperature rises. Or, conversely, the colder the 
temperature, the greater amount of 02(g) would be dissolved. This is an important environmental concept 
in understanding the effect of the increase in the temperature in the world's oceans and its effect on the 
oceanic life. 

The Effect of Pressure 

The second factor, pressure, affects the solubility of a gas in a liquid but never a solid dissolving in a liquid. 
When pressure is applied to a gas that is above the surface of a solvent, the gas will move into the solvent 
and occupy some of the spaces between the particles of the solvent. A good example is carbonated soda. 
Pressure is applied to force the CO2 molecules into the soda. The opposite is also true. When the gas 
pressure is decreased, the solubility of that gas is decreased. One example you can think about is when you 
open a can of soda pop, or any carbonated beverage, for that matter. When you hear the "pop" as a can 
or bottle of soda pop is opened, the pressure in the soda is lowered and the gas immediately starts leaving 
the solution. The carbon dioxide stored in the soda is released and you see the fizzing on the surface of 
the liquid. 

This gas pressure factor is expressed in Henry's Law. Henry's Law states that at a given temperature 
the solubility of a gas in a liquid is proportional to the pressure of the gas above the liquid. An example of 
Henry's Law occurs in SCUBA diving. As a person dives into deep water, the pressure increases and more 
gases are dissolved into the blood. While ascending from a deep water dive, the diver needs to return to 
the surface of the water at a very slow rate since all of the dissolved gases have to come out of the blood 
very slowly. If for some reason a person ascends too quickly, a medical emergency may occur since the 
gases will come out of the blood too quickly. This is called having the "bends". 

Increased Surface Area Increases Rate of Dissolving 

Another factor that affects the rate of solubility is the surface area of a solid. If we were to increase the 
surface area of a solid then it would have been broken into smaller pieces. We would do this to increase 



www.ckl2.org 



436 



how quickly the solute would dissolve in solution. If you were to dissolve sugar in water, a sugar cube will 
dissolve slower than an equal amount of tiny pieces of sugar crystals. The combined surface area of all of 
the sugar crystals have a much greater surface area than the one sugar cube and therefore will have more 
contact with the water molecules. This allows the sugar crystals to dissolve much more quickly. Try it on 
your own and see! What other examples can you think of? 

If you were working in a lab, you might be asked to make a solution of copper(II) sulfate. Copper(II) sulfate 
comes in several forms: one being a large beautiful blue crystal and the other ground-up blue crystals. 
When you set equal amounts of each of these is test tubes filled with 10 mL of water, you will notice after 
5 minutes that more of the ground-up crystal will have dissolved (and the solution will be a darker blue) 
than the test tube with the large crystal. Or, you can take two samples of the ground-up crystal and put 
into separate test tubes. This time stopper one of the test tubes and carefully shake it while you let the 
other test tube simple sit still. Again you are increasing the surface area by increasing the how much of the 
ground-up crystal will come in contact with the water. The result will still be the same as before: the test 
tube with the greater surface area will go into solution at a faster rate. Even though maximum solubility 
is achieved more quickly with greater surface area, the concentration of the solute at maximum solubility 
will be exactly the same. (See Figures 15.6 and 15.7). 




Figure 15.6: Crystalline copper (II) sulfate. 




Figure 15.7: Powdered copper (II) sulfate. 



Lesson Summary 



Temperature affects the solubility of both gases and solids. With solids, generally the solubility 
increases with increasing temperature. With gases, the solubility tends to decrease with increasing 
temperature. 



437 



www.ckl2.org 



• Pressure only affects the solubility of gases. Henry's Law states that at a given temperature the 
solubility of a gas in a liquid is proportional to the pressure of that gas. 

• Increasing the surface area increases the rate of solubility of a solid because a larger number of 
molecules of the greater surface area have contact with the solvent. 

Review Questions 

1. What are the factors that affect solubility? 

2. What is Henry's Law? 

3. Is it ever possible to have ionic solids decrease solubility with increasing temperature? 

4. What factor would affect the solubility of sodium sulfate? 

(a) i, ii, and iii 

(b) i and ii 

(c) i and iii 

(d) ii and iii 

i. temperature 
ii. pressure 
iii. surface area 

5. What factor would affect the solubility of methane? 

(a) i, ii, and iii 

(b) i and ii 

(c) i and iii 

(d) ii and iii 

i. temperature 
ii. pressure 
iii. surface area 

6. If you crush a cube a sugar before putting it in your cup of coffee, how have you affected its solubility? 

(a) Crushing it has really no affect on solubility because we have not heated it at all. 

(b) Crushing it has increased the surface area so it speeds up the dissolving process but doesn't 
change maximum solubility. 

(c) Crushing it has really no affect on solubility because we have not stirred it at all. 

(d) Crushing it has increased the surface area so it increases the maximum solubility. 

7. Why do people add chlorine to their swimming pools on a hot day? 

8. Explain why crushed table salt at room temperature dissolves faster than rock salt. 

9. Under which of the following sets of conditions would the solubility of CO2C?) be lowest? The pressure 
given is the pressure of C 02(g) above the solution. 

(a) 5.0 atm and 75° C 

(b) 1.0 atm and 75° C 

(c) 5.0 atm and 25°C 

(d) 1.0 atm and 25°C 

(e) 3.0 atm and 25°C 

10. An aqueous solution of KCl is heat from 15°C to 85°C Which of the following properties of the 
solution remain the same? 

(a) i only 

(b) iii only 

(c) i and ii only 

www.ckl2.org 438 



(d) ii and iii only 

(e) i, ii, and iii 

i. molality 
ii. molarity 
iii. density 

Vocabulary 

Henry's Law At a given temperature the solubility of a gas in a liquid is proportional to the pressure 
of that gas. 

15.7 Colligative Properties 

Lesson Objectives 

• Describe vapor pressure lowering. 

• Define boiling point elevation and freezing point depression. 

• Describe what happens to the boiling points and freezing points when a solute is added to a solvent. 

• Describe the importance of the van't Hoff factor. 

• Calculate the boiling point elevation for electrolyte and non-electrolyte solutions. 

• Calculate the freezing point depression for electrolyte and non-electrolyte solutions. 

Pure solvents and solutions of solutes in solvents vary in their boiling points and freezing points. People 
who live in colder climates have seen the trucks put salt on the roads when snow or ice is forecast. Why 
do they do that? When planes fly in cold weather, the planes need to be de-iced before liftoff. Why is that 
done? In this lesson you will learn why these events occur. You will also learn to calculate exactly how 
much of an effect a specific solute can have on the boiling point or freezing point of a solution. 

Vapor Pressure Lowering 

In the chapter on liquids, you studied the concept of vapor pressure for liquids. An enclosed liquid will 
reach a vapor pressure equilibrium with its vapor in the space above the liquid and this vapor pressure 
depends on the temperature of the liquid. This vapor pressure equilibrium is reached when the rate of 
evaporation and the rate of condensation become equal. Raising the temperature increases the rate of 
evaporation and therefore, increases the vapor pressure of the liquid. When the temperature of the liquid 
becomes high enough for the vapor pressure to equal the surrounding pressure, the liquid will boil. When 
the surrounding pressure is 1.00 atm, the boiling point is called the "normal" boiling point. 

A pure liquid solvent, at normal atmospheric pressure, cannot have its temperature raised above the normal 
boiling point because at the normal boiling point, all of the liquid will vaporize before the temperature 
will increase. 

Adding a solute to a solvent lowers the vapor pressure of the solvent. There are two suggested explanations 
for why the addition of a solute lowers the vapor pressure of a solution. Since they seem equally valid, 
both will be presented here. 

Remember that only the molecules on the surface of a liquid are able to evaporate. (See Figure 15.8). 

In a pure solvent, all the molecules at the surface are solvent molecules. Therefore, the entire surface area 
is available for evaporation and the forces to be overcome are the attractive forces between the solvent 

439 www.ckl2.org 



Pure Solvent Solution 



Figure 15.8: Pure solvent particles versus intermixed solute particles. 



molecules. One of the explanations says that in a solution, some of the surface molecules are solute 
molecules and since these solute molecules take up some of the surface area, less surface area is available 
for evaporation. Therefore, the rate of evaporation of the solvent will be lower and so the vapor pressure 
will be lower at the same temperature. The other explanation says that the attractive forces between the 
solvent molecules and the solute molecules are greater than the attractive forces between solvent molecules 
and therefore, the solvent molecules will not evaporate at as high a rate. Once again vapor pressure will 
be lowered. Both explanations start with the same premises and end with the same result so there doesn't 
seem to be a reason to choose between them. 

The amount of vapor pressure lowering is related to the molal concentration of the solute. Suppose a 
pure solvent has a vapor pressure of 20. mm of Hg at a given temperature. Suppose further that sufficient 
non-electrolyte solvent is dissolved in the solvent to make a 0.20 m solution and that the vapor pressure 
of the solution at the same temperature is 18 mm of Hg. Then, if enough solute is dissolved to make the 
solution 0.40 m, the vapor pressure of the more concentration solution will be 16 mm of Hg. There is a 
direct mathematical relationship between the molality of the solution and the vapor pressure lowering. 

Experiments with this phenomenon demonstrate that as long as the solute is a non-electrolyte, the effect 
is the same regardless of what solute is used. The effect is related only to the number of particles of solute, 
not the chemical composition of the solute. That is the definition of the term colligative properties; 
properties that are due only to the number of particles in solution and not related to the chemical properties 
of the solute. Any polar, non-ionic solid used to make a 0.20 m solution would cause the exact same vapor 
pressure lowering. 



Boiling Point Elevation 

The boiling point of a liquid occurs when the vapor pressure above the surface of the liquid equals the 
surrounding pressure. We know that water boils at 100°C at 1 atm but a solution of salt water does not. 
When table salt is added to water the resulting solution has a higher boiling point because the salt causes 
the vapor pressure of the solvent to lower. The ions form an attraction with the solvent particles which then 
prevent the water molecules from going into the vapor phase. Pure water at 100°C has a vapor pressure 
of 1.00 atm and since the surrounding pressure is 1.00 atm, the pure water boils at this temperature. If 
some quantity of table salt is added to the water, the vapor pressure of the solution at 100°C might be 
only 0.98 atm. Therefore, the salt water solution will not boil at 100°C In order to cause the salt water 
solution to boil, the temperature must be raised above 100°C in order to get the vapor pressure up to 
1.00 atm and allow the solution to boil. This is true for any solute added to a solvent; the boiling point 
of the solution will be higher than the boiling point of the pure solvent (without the solute). The boiling 
point elevation is the difference in the boiling points of the pure solvent and the solution. 

The boiling point elevation due to the presence of a solute is also a colligative property. That is, the 
amount of change in the boiling point is related to number of particles of solute in a solution and is not 
related to chemical composition of the solute. A 0.20 m solution of table salt and a 0.20 m solution of 
hydrochloric acid would have the same effect on the boiling point. 

www.ckl2.org 440 



Freezing Point Depression 

The effect of adding a solute to a solvent has the opposite effect on the freezing points of a solution as it 
does on the boiling point. A solution will have a lower freezing point than a pure solvent. The freezing 
point is the temperature at which the liquid changes to a solid. At a given temperature, if a substance is 
added to a solvent (such as water), the solute-solvent interactions prevent the solvent from going into the 
solid phase. The solute-solvent interactions require the temperature to decrease further in order to solidify 
the solution. A common example is found when salt is used on icy roadways. Here the salt is put on the 
roads so that the water on the roads will not freeze at the normal 0°C but at a lower temperature, under 
ideal conditions to as low as — 9°C. The de-icing of planes is another common example of freezing point 
depression in action. A number of solutions are used but commonly a solution such as ethylene glycol, or 
a less toxic monopropylene glycol, is used to de-ice an aircraft. The aircrafts are sprayed with the solution 
in weather conditions suspected of dropping below the freezing point. The freezing point depression is 
the difference in the freezing points of the solution from the pure solvent. 

Mathematics of Boiling Point and Freezing Point Changes 

The boiling point of a solution is higher than the boiling point of a pure solvent and the freezing point of 
a solution is lower than the freezing point of a pure solvent. However, the amount to which the boiling 
point increases or the freezing point decreases depends on the amount solute that is added to the solvent. 
A mathematical equation is used to calculate the difference in difference of the boiling point elevation or 
the freezing point depression. The boiling point elevation is the difference between the boiling points of 
the pure solvent and the solution. Remember the solution has a higher boiling point so to find the boiling 
point elevation you would subtract the boiling point of the solvent from the boiling point of the solution. 
For example, the boiling point of pure water at 1.0 atm is 100. °C while the boiling point of a 2% salt 
water solution is about 102°C Therefore, the boiling point elevation would be 2°C The freezing point 
depression is the difference in the freezing points of the solution from the pure solvent. Remember that 
the solution has a lower freezing point so the freezing point depression is found by subtracting the freezing 
point of the solution from the freezing point of the pure solvent. 

Both the boiling point elevation and the freezing point depression are related to the molality of the solutions. 
Looking at the formulas for the boiling point elevation and freezing point depression, we can see similarities 
between the two. 

ATb = Kbin 

ATf = Kfm 

where: 

ATt, = T(pure solution)-r(pure solvent) 
ATf = T(pure solvent)-r(pure solution) 
Kb = boiling point elevation constant 
Kf = freezing point depression constant 
m = molality of the solution 
m = molality of the solution 

The boiling point constants and freezing point constants are different for every solvent and are determined 
experimentally in the lab. You can find these constants for hundreds of solvents listed in data reference 
publications for chemistry and physics. 

441 www.ckl2.org 



Sample question 1: Antifreeze is used in automobile radiators to keep the coolant from freezing. In 
geographical areas where winter temperatures go below the freezing point of water, using pure water as 
the coolant could allow the water to freeze. Since water expands when it freezes, freezing coolant could 
crack engine blocks, radiators, and coolant lines. The main component in antifreeze is ethylene glycol, 
C2Hi{OH)2- If the addition of an unknown amount of ethylene glycol to 150 g of water dropped the 
freezing point of the solution by — 1.86°C, what mass of ethylene glycol was used? The freezing point 
constant, Kf for water is -1.86° C/m. 

Solution: 

ATf = Kfin 

At f -1.86°C 

m = — - = — - — = 1.00 m 

K f -1.86°C/m 



mols solute 



m 



kg solvent 

moles solute = (molality) ( kg solvent) = (1.00 mol/kg) (0.150 kg) = 0.150 mol 
mass C2Hi{OH)2 = ( mols)( molar mass) = (0.150 mol)(62.1 g/mol) 
mass C 2 H 4 {OH) 2 = 9.3 g 

Therefore 9.3 g of ethylene glycol would have been added to the 150. g of water to lower the freezing point 
by 1.86°C. 

One factor that must be considered when using these formulas is to first determine whether or not the 
solution is an electrolyte or a non-electrolyte. 

Recall that an electrolyte solution is one where the solute will separate into ions and is thus able to conduct 
electricity (hence the term electrolyte). If, for example, you have a solution of sodium chloride, the sodium 
chloride separates into sodium ions and chloride ions. A non-electrolyte solution does not separate into 
ions and therefore does not conduct electricity. For example a sugar /water solution stays as sugar + water 
with the sugar molecules staying as molecules. 

Remember that colligative properties are due to the number of solute particles in the solution. Adding 
10 molecules of sugar to a solvent will produce 10 solute particles in the solution. When the solute is an 
electrolyte, such as NaCl however, adding 10 molecules of solute to the solution will produce 20 ions (solute 
particles) in the solution. Therefore, adding enough NaCl solute to a solvent to produce a 0.20 m solution 
will have twice the effect of adding enough sugar to a solvent to produce a 0.20 m solution. It is the 
number of solute particles in the solution that control the colligative properties. 

To add this "electrolyte" factor into the formulas above, we use a concept called the van't Hoff factor, 
symbolized by the letter i. The van't Hoff factor is the number of particles that the solute will dissociate 
into upon mixing with the solvent. For example, sodium chloride, NaCl, will dissociate into two ions so 
the van't Hoff factor for NaCl is i = 2, for lithium nitrate, LiNOz, i = 2, and for calcium chloride, CaCh, 
i = 3. 

We can now rewrite our colligative properties formulas and include the van't Hoff factor. 

AT i, — iKbtn 
ATf = iKfin 

where: 

www.ckl2.org 442 



ATb = T(pure solvent )-T (pure solution) 
ATf = T(pure solvent )-T (pure solution) 
i = van't Hoff factor 
Kb = boiling point elevation constant 
Kf = freezing point depression constant 
m = molality of the solution 

Why can we use this formula for both electrolyte and non-electrolyte solutions? Since the van't Hoff factor 
for non-electrolytes is always 1 (because they do not dissociate), i is always equal to 1. 

Sample question 2: A solution of 10.0 g of sodium chloride is added to 100.0 g of water in an attempt to 
elevate the boiling point. What is the boiling point of the solution? 

Solution: 

ATb — iKbtn 

mols NaCl = f™ = 1 °-° / S - = 0.171 mols 
molar mass 58.5 g/mol 

, , mols solute 0.171 mol „ „ _, . , „, . 

molality = = = 1.71 m For NaCLj = 2 (NaCl -> Na + + CV) 

kg solvent 0.100 kg v ' 

K b ( water) = 0.52° C/m 

ATb = iKb m 

AT h = (2)(0.52° C/m)(1.71 m) 

AT,, = 1.78°C 
^(solution) = 7^ (pure solvent) + ATb 
resolution) = 100°C + 1.78°C = 101.78°C. 

Therefore the boiling point of the solution of 10 g of NaCl in 100 g of water is 102°C 

Lesson Summary 

• Boiling points of solutions are higher that the boiling points of the pure solvents. Freezing points 
of solutions are lower than the freezing points of the pure solvents. Boiling point elevation is the 
difference between the boiling points of the pure solvent and the solution. The boiling point elevation 
can be calculated using the formula ATb = Kb m, where ATb is the boiling point elevation, Kb is the 
boiling point elevation constant, and m is molality. 

• Freezing point depression is the difference between the freezing points of the solution and the pure 
solvent. The freezing point depression can be calculated using the formula ATf = Kf m, where ATf 
is the freezing point depression, Kf is the freezing point depression constant, and m is molality. 

• For electrolyte solutions, the van't Hoff factor is added to account for the number of ions that the 
solute will dissociate into in solution. For non-electrolyte solutions, the van't Hoff factor = 1. The 
equations change to account for this factor (ATb = Kb m becomes ATb = 1Kb m and ATf = Kf m 
becomes ATf = iKf m, where i is the number of particles each solute molecule produces in solution. 

Review Questions 

1. What must be measured in order to determine the freezing point depression? 

443 www.ckl2.org 



2. From a list of solutions with similar molalities, how could you quickly determine which would have 
the highest boiling point? 

3. Why would table salt not be a good solution to use when deicing a plane? 

4. When a solute is added to a solution: 

(a) i and hi are true 

(b) i and iv are true 

(c) ii and iii are true 

(d) ii and iv are true 

i. the boiling point increases. 

ii. the boiling point decreases, 

iii. the freezing point increases, 

iv. the freezing point decreases. 

5. If 25.0 g of sucrose (C12H22O11) is added to 500. g of water, the boiling point is increased by what 
amount? (K b ( water) = 0.52° C/m) 

(a) 0.076° 

(b) 0.025° 

(c) 26° 

(d) None of these 

6. The solubility of seawater (an aqueous solution of NaCl) is approximately 0.50 m. Calculate the 
freezing point of seawater. (Kf ( water) = 1.86° C/m) 

(a) -0.93° 

(b) 0.93° 

(c) 1.86° 

(d) -1.86° 

7. Determine which of the following solutions would have the lowest freezing point. 

(a) 15 g of ammonium nitrate in 100. g of water. 

(b) 50. g of glucose in 100. g of water. 

(c) 35 g of calcium chloride in 150. g of water. 

8. A 135.0 g sample of an unknown nonelectrolyte compound is dissolved in 725 g of water. The 
boiling point of the resulting solution was found to be 106.02°C What is the molecular weight of the 
unknown compound? 

9. What is the van't Hoff factor for each of the following: 

(a) MgCl 2 

(b) Ammonium sulfate 

(c) CH 3 OH 

(d) Potassium chloride 

(e) KCH 3 COO 

10. Calcium chloride is known to melt ice faster than sodium chloride but is not used on roads because 
the salt itself attracts water. If 15 g of CaCli was added to 250 g of water, what would be the effect 
on the freezing point of the solution? (Kf( water) = -1.86° C/m) 

Vocabulary 

boiling point elevation The difference in the boiling points of the pure solvent from the solution. 

freezing point depression The difference in the freezing points of the solution from the pure solvent. 
www.ckl2.org 444 



Van't Hoff factor The number of particles that the solute will dissociate into upon mixing with the 
solvent 

15.8 Colloids 

Lesson Objectives 

• Define colloids and suspensions. 

• Compare solutions, colloids, and suspensions. 

• Characterize solutions as suspensions, colloids, or solutions. 

• Name some common examples of colloids. 

Introduction 

In this final section regarding solution chemistry, we will take a look at colloids and suspensions; two 
different types of mixtures that are not classified as true solutions. Following this, a look at the different 
types of colloids in terms of the solute/solvent combinations will be examined to see how we have many 
examples in our everyday lives. Do you know that the majority of our encounters with mixtures are actually 
colloids? Think of that the next time you drink a glass of milk, put whipping cream on your dessert, look 
up at the clouds, or add some butter to your toast! 

Comparison of Suspensions, Colloids and Solutions 

We learned early in the chapter that a solution is a mixture of substances in such a way that the final 
product has the same composition throughout. Remember the example of vinegar that is 5%, by mass, 
acetic acid in water. This clear liquid is a solution since light easily passes through it and it never separates. 

On the other hand, colloids are mixtures in which the size of the particles is between 1 X 10 3 pm and 
1 X 10 6 pm. In meters, these sizes translate to 1 X 10~ 9 m to 1 X 1CT 6 m; a small grain of sand has a diameter 
of 2 x 1CT 5 m. A common example for a colloid is milk. One way to tell that milk is a colloid is to test 
it using the Tyndall Effect. The Tyndall effect involves shining a light through the mixture: when the 
light is not allowed to pass through the mixture, that is, the light is scattered, the mixture is considered 
a colloid. This is why milk appears "cloudy" - or what we think of as "milky". In contrast, when light is 
passed through a true solution, the particles in solution are so small [atoms, ions, small molecules] that 
they do not obstruct the light. However, when light is passed through a colloid, since the particles are 
larger, they will act as an obstruction to the light and the light is scattered. However, these particles, 
while able to scatter light, are still small enough so that they do not settle out of solution. 

In contrast, suspensions are mixtures where the particles settle to the bottom of the container which 
means that the particles in a suspension are large enough so that gravity pulls them out of solution. With 
suspensions, filtration is usually able to be used to separate the excess particles from the solution. A 
common example of a suspension is muddy water. If you had a beaker of water and added a handful of 
fine dirt, even if you stirred it (making a colloid type solution), letting it stand, dirt would settle to the 
bottom. 

Many Common Products Are Colloids 

It is amazing just how common colloids are to us in our everyday lives. In our earlier lessons we discussed 
solutes and solvents and what types of solutions formed as a result. In this final section of this final lesson, 

445 www.ckl2.org 



let's do the same process. Looking at Table 15.5, we see that some common colloid products that are 
formed when different phase solutes and solvents are mixed. 

Table 15.5: 



Solid Solvent 



Liquid Solvent 



Gas Solvent 



Solid Solute Ruby, brass, steel Butter, cheese, mus- Marshmallow 

tard, Jell-0 
Liquid Solute Paint, milk of magnesia Milk, Mayonnaise, Whipping Cream, Shav- 

Creams (i.e. face ing Cream 

creams) 
Gas Solute Airborne viruses, car Fog, Smoke, Clouds, Aerosol Sprays 

exhaust, smoke 



Lesson Summary 

Colloids are mixtures in which the size of the particles is between 1 x 10 3 pm and 1 x 10 6 pm. Suspensions 
are mixtures in which the particles are large enough so that they settle to the bottom of the container and 
can be filtered using filter paper. The Tyndall effect involves shining a light through the mixture, if the 
light scatters, the mixture is a colloid or a suspension. Common examples of colloids include milk, butter, 
Jell-O, and clouds. 



Review Questions 



1. Distinguish between a solution, a colloid, and a suspension. 

2. What is one true way to tell you have a colloid solution? 

3. Why do you think there is no example of a gas - gas colloid? 

4. Which is an example of a colloid? 

(a) air 

(b) brass 

(c) milk 

(d) none of these 

5. Which is not an example of a colloid? 

(a) human body 

(b) mayonnaise 

(c) mustard 

(d) cloud 

6. The biggest difference between a colloid and a suspension is that: 

(a) In colloids, the solute is permanently dissolved in the solvent. 

(b) In colloids the particles eventually settle to the bottom. 

(c) In suspensions the particles eventually settle to the bottom. 

(d) None of these are correct 

7. Karen was working in the lab with an unknown solution. She noticed that there was no precipitate 
in the bottom of the beaker even after it had been on the lab bench for several days. She tested it 
with a light and saw that light scattered as it passed through the solution. Karen concluded that the 
liquid was what type of a mixture? 



www.ckl2.org 



446 



(a) colloid 

(b) suspension 

(c) homogeneous 

(d) heterogeneous 

8. What are two good common examples of colloids? 

Vocabulary 

colloid Mixtures where the size of the particles is between 1 x 10 3 pm and 1 x 10 6 pm (i.e., milk). 

suspension Mixtures where the particles settles to the bottom of the container and can be separated by 
filtration. 

Tyndall Effect Involves shining a light through the mixture, if the light scatters, the mixture is a colloid. 



15.9 Separating Mixtures 

Lesson Objectives 



• The student will describe differences between the physical properties of pure substances and solutions. 

• The student will list and describe methods of separation for mixtures. 

• The student will explain the principles involved in chromatographic separation. 

• The student will identify the mobile and stationary phases in a chromatography design. 

• Given appropriate data, the student will calculate Rf values. 

Introduction 

Mixtures occur very commonly in chemistry. When a new substance is synthesized, for example, the new 
substance usually must be separated from various side-products, catalysts, and any excess reagent still 
present. When a substance must be isolated from a natural biological source, the substance of interest is 
generally found in a very complex mixture with many other substances, all of which must be removed. 
Chemists have developed a series of standard methods for the separation of mixtures. (The separation of 
mixtures into their constituent substances defines an entire sub-field of chemistry referred to as separation 
science. 

Differing Solubilities 

Mixtures of solids may often be separated on the basis of differing solubilities of the solids. If one 
of the components of the mixture is soluble in water while the other components are insoluble in water, 
the water-soluble component can be removed from the mixture by dissolving the mixture in water and 
filtering the mixture through filter paper. The component dissolved in water will pass through the filter 
while the undissolved solids will be caught in the filter. The solubility of substances is greatly influenced 
by temperature. By controlling the temperature at which solution occurs or at which the filtration is 
performed, it may be possible to separate the components. Most commonly, a sample is added to water 
and heated to boiling. The hot sample is then filtered to remove completely insoluble substances. The 
sample is then cooled to room temperature or below which causes crystallization of those substances whose 

447 www.ckl2.org 



solubilities are very temperature dependent. These crystals can then be separated by another filtration and 
the filtrate (the stuff that went through the filter) will then contain only those substances whose solubilities 
are not as temperature dependent. 

Distillation 

Homogeneous solutions are most commonly separated by distillation. (See Figure 15.9). In general, 
distillation involves heating a liquid to its boiling point, then collecting, cooling, and condensing the 
vapor produced into a separate container. For example, salt water can be desalinated by boiling off and 
condensing the water. 

In solutions of non-volatile solid solutes in liquid solvent, when the solution is boiled, only the solvent boils 
off, all the solid remains in the solution. The vapor passing off is pure solvent. As the solvent passes off 
and the all the solute remains behind, the same amount of solute is now dissolved in less solvent and so 
the concentration increases. This results in the boiling point of the solution increasing. As a solution boils, 
increased temperature is necessary to keep the solution boiling because its boiling point has increased. This 
is a quick method of determining if a liquid is a pure substance or a solution; start it boiling and if it 
continues to boil at the same temperature, it is a pure substance whereas if its boiling point increases, it is 
a solution. 

For a mixture of liquids in which several components of the mixture are likely to be volatile (easily vapor- 
ized), the separation is not easy. If the components of the mixture differ reasonably in their boiling points, 
it may be possible to separate the mixture simply by monitoring the temperature of the vapor produced 
as the mixture is heated. Liquid components of a mixture will each boil in turn as the temperature is 
gradually increased, with a sharp rise in the temperature of the vapor being distilled indicating when a 
new component of the mixture has begun to boil. By changing the receiving flask at the correct moment, 
a separation can be accomplished. This process is known as fractional distillation. 




Figure 15.9: Distillation Apparatus 



Chromatography 

The word chromatography means color- writing. The name was chosen around 1900 when the method was 
first used to separate colored components from plant leaves. Chromatography in its various forms is perhaps 



www.ckl2.org 



448 



the most important known method of chemical analysis of mixtures. Paper and thin-layer chromatography 
are simple techniques that can be used to separate mixtures into the individual components. The methods 
are very similar in operation and principle and differ primarily in the medium used. 

Paper chromatography uses ordinary filter paper as the medium upon which the mixture to be separated is 
applied. Thin-layer chromatography (abbreviated TLC) uses a thin coating of aluminum oxide or silicagel 
on a glass microscope slide or plastic sheet to which the mixture is applied. A single drop of the unknown 
mixture to be separated is applied about half an inch from the end of a strip of filter paper or TCL slide. 
The filter paper or TLC slide is then placed in a shallow layer of solvent in a jar or beaker. Since filter 
paper or the TLC slide coating is permeable to liquids, the solvent begins rising up the paper by capillary 
action. 

As the solvent rises to the level of the mixture spot, various effects can occur, depending on the constituents 
of the spot. Those components of the spot that are completely soluble in the solvent will be swept along 
with the solvent front as it continues to rise. Those components that are not at all soluble will be left 
behind at the original location of the spot. Most components of the mixture will move up the paper or 
slide at an intermediate speed somewhat less than the solvent front speed but not remaining at the original 
spot either. In this way, the original spot of mixture is spread out into a series of spots or bands, with each 
spot representing one single component of the mixture. The separation of a mixture by chromatography 
is not only a function of the solubility in the solvent used. The filter paper or TLC coating consists of 
molecules that may interact with the molecules of mixture as they are carried up the medium. The primary 
interaction between the mixture components and the medium is due to the polarity of the components and 
that of the medium. Each component of the mixture is likely to interact with the medium to a different 
extent thus slowing the components of the mixture differentially depending on the level of interaction. (See 
Figure 15.10). 



Solvent front 



Intermediate Spots 



Origbal Spot 



Line of Origb 




Figure 15.10: Paper chromatography strip 

A quantitative basis is added to the chromatography analysis using a mathematical function call the 
retention factor. The retention factor, Rf, is defined as 



Rt 



distance travelled by spot 
distance travelled by solvent front 



Rf is the ratio of the distance a substance moves up the stationary phase to the distance the solvent have 
moved. The retention factor depends on what solvent is used and on the specific composition of the filter 
paper or slide coating used. The Rf value is characteristic of a substance when the same solvent and the 
same type of stationary phase is used. Therefore, a set of known substances can be analyzed at the same 



449 



www.ckl2.org 



time under the same conditions and if the same Rf value is produced for a known and unknown, it is a 
step toward identifying the unknown. 

In the case shown at right, the Rf for the green spot is 

2.7 cm 



R 



f 



5.7 cm 



0.47 



and for the yellow spot 



1.8 cm 

R f = — = 0.32 

5.7 cm 



5.7 cm 




2.7 cm 



1.8 cm 



cm 



Paper chromatography and TLC are only two examples of many different chromatographic methods. Mix- 
tures of gases are commonly separated by gas chromatography. In this method, a mixture of liquids are 
vaporized and passed through a long tube of solid absorbent material. A carrier gas, usually helium, is 
used to carry the mixture of gases through the tube. As with paper chromatography, the components of 
the mixture will have different solubilities and different attractions for the solid absorbent. Separation of 
the components occurs as the mixture moves through the tube. The individual components exit the tube 
one by one and can be collected. 

Another form of chromatography is column chromatography. In this form, a vertical column is filled with 
solid absorbent, the mixture is poured in at the top, and a carrier solvent is added. As the mixture flows 
down the column, the components are separated, again, by differing solubilities in the carrier solvent and 
different absorbencies to the solid packing. As the liquid drips out the bottom of the column, components 
of the solution will exit at different times and can be collected. 



Lesson Summary 

• Mixtures of solids may be separated by differing solubilities of the solids. 

• Components of a solution composed of a non- volatile solid solute and a liquid solvent can be separated 
by distillation. 

• Mixtures of liquids with reasonably different boiling points can also be separated by distillation. 



www.ckl2.org 



450 



• Solutions with several components can be separated by paper or thin-layer chromatography. 

• Gas chromatography and column chromatography are also used to separate the components of a 
solution. 

Review Questions 

1. In a paper chromatography experiment to separate the various pigments in chlorophyll, a mixture of 
water and ethanol was used as the solvent. What is the stationary phase in this separation? 

2. Do you think that paper chromatography or TLC would be useful for separating a very large quantity 
of a mixture? Explain why or why not. 

3. If the mobile phase in a chromatographic experiment moved 15.0 cm and one of the compounds in 
the mixture moved 12.7 cm, what is the Rf value for this compound? 

4. If the stationary phase in a paper chromatography experiment was very polar and the solvent was 
moderately polar, would the polar components in the mixture be closer to the bottom of the paper 
or toward the top of the paper? 

Vocabulary 

distillation The evaporation and subsequent collection of a liquid by condensation as a means of purifi- 
cation. 

fractional distillation This is a special type of distillation used to separate a mixture of liquids using 
their differences in boiling points. 

chromatography Any of various techniques for the separation of complex mixtures that rely on the 
differential affinities of substances for a mobile solvent and a stationary medium through which they 
pass. 



Image Sources 



(i 

(2 
(3 
(4 
(5 

(6 

(7 

(8 

(9 

(10 



Powdered copper (II) sulfate.. Public Domain. 

Therese Forsythe. A solubility graph for common ions and ammonia.. CC-BY-SA. 
CK-12 Foundation. Reading a solubility graph.. CC-BY-SA. 

Crystalline copper (II) sulfate.. Public Domain. 
H Padleckas. Distillation Apparatus. GNU Free Documentation License. 
CK-12 Foundation. Reading a solubility graph.. CC-BY-SA. 
Richard Parsons. The hydration of ions in a polar solvent.. CC-BY-SA. 
Richard Parsons. Paper chromatography strip. CC-BY-SA. 
Richard Parsons. . CC-BY-SA. 
CK-12 Foundation. . CC-BY-SA. 

451 www.ckl2.org 



Chapter 16 



Ions in Solution 



16.1 Ionic Solutions 

Lesson Objectives 

• Describe electrostatic attraction. 

• Explain how ionic solids attract water molecules when they dissolve in water. 

• Explain the difference between physical changes and chemical changes. 

• Define electrolyte solutions and be able to identify electrolytes. 

Introduction 

Ionic solids are a particular type of substance. They form when metal ions combine with nonmetal ions, 
which come about through the transfer of electrons. Because of this transfer and the distinct charges that 
result when the metals and nonmetal ions form, ionic solids have properties that are unique to themselves. 
In this lesson we will begin to look at the forces that exist within the ionic solids and what happens when 
these solids dissolve in water. There are some new terms in this unit and a few reviewed terms that you 
have seen in previous units that will become clearer in this chapter. 

Ions in Solids are Held With Electrostatic Attraction 

In an earlier chapter you learned that metals form positive ions by losing electrons and nonmetals form 
negative ions by gaining electrons. When solids form from a metal atom donating an electron (thus forming 
a positive cation) to a nonmetal atom (thus forming a negative anion) , the ions in the solid are held together 
by the attraction of these oppositely charges particles. The attraction of oppositely charged particles is 
called electrostatic attraction. For example: 

Na + ionization energy — » Na + + e~ 

I + e~ — » I + energy of electron affinity 

Na + + I — » Nal + energy 

(Nal is held together by electrostatic attraction) 

In the world of chemical bonding, electrostatic attraction is quite strong and therefore compounds with 

www.ckl2.org 452 



this type of bonding have high melting points and boiling points. Take a look at Table 16.1 and see how 
the melting points and boiling points change for ionic solids. 

Table 16.1: 



Compound 



Melting Point (°C) 



Boiling Point (°C) 



LiCl 

NaCl 

KCl 

MgCl 2 

CaCh 



613 
801 

772 
714 

772 



1360 
1413 
1500 
1412 
1935 



All of the compounds in the table are bonded due to the attraction of the oppositely charged ions. Look 
at the next examples to see lithium, sodium, and potassium chlorides. 



Li — > Li + + e- 
Cl +e- — > Cl- 
Li + +C|-— ► LiCl 




Na — > Na + + e 
CI + e - — » CI - 
Na + +CI- — > NaCl 




NaCl 



(s) 



453 



www.ckl2.org 



K — ► K + + e 
CI +e~ — *CI~ 
K + + CI - — ► KCI 




KCI 



(s) 



To review from a previous chapter, ionic compounds do not form molecules. The empirical formula repre- 
sents the lowest whole number ratio of the ions involved in the compound. When we interact with these 
substance in our environment, we detect that they are crystalline structures. As shown in the diagrams, 
ionic compounds are crystals that are held together by electrostatic attraction. 

Sample question 1: Which compounds would contain an electrostatic attraction as bonds between ions? 

(a) MgCh 

(b) Al 2 3 

(c) CH 4 

Solution: 

(a) and (b) would contain electrostatic attraction because they are both ionic but (c) is not. It does not 
form bonds by the transfer of electrons but rather by sharing electrons and therefore does not have ions 
for electrostatic attraction. 



Ions Are Attracted to the Polar Water Molecule 

Since ionic compounds can dissolve in polar solutions, specifically water, we can extend this concept to say 
that ions themselves are attracted to the water molecules because the ions of the ionic solid are attracted 
to the polar water molecule. When you dissolve table salt in a cup of water, the table salt dissociates into 
sodium ions and chloride ions (Equation 1). 



NaCl(s) — > Na + ( aq -j + CI r aq \ (Equation 1) 



The sodium ions then get attracted to the negative ends of the water molecule and the chloride ions get 
attracted to the positive end of the water molecule. The process of water molecules attaching to ions is 
called hydration. Look at the Figure ??. 

www.ckl2.org 454 







Sodium Chloride 
Crystal 



Na + 

Na 

Na + CI- Na + 
Na + 




Hydrated Na + ion 





Na + Cl- 
Na + 
Na + CI- Na + 





Hydrated Ch ion 



8 « * 

CI- Na + Cl| 



The same is true for any ionic compound dissolving in water. The ionic compound will separate into the 
positive and negative ions and the positive ion will be attracted to the negative end of the water molecules 
(oxygen) while the negative ion will be attracted to the positive end of the water molecules (hydrogen) . 

Dissociation is a Physical Change 

Matter can go through both chemical changes and physical changes. Chemical changes are ones that 
occur with the chemical bonding and a new substance or substances are formed. For example, if you add 
a piece of lead to a solution of silver nitrate, silver precipitates out and lead nitrate is formed (Equation 

2). 

Pb (s) + 2 AgN0 3{aq) -> 2 Ag (s) + Pb(N0 3 ) 2(aq) (Equation 2) 

Of importance to us in this unit are physical changes. Physical changes are those that occur in the 
physical state of the substance but do not affect the identity of the substance. For example, grinding a 
sugar cube is a physical change or dissolving table salt in water; the sugar is still sugar and the table salt 
is still table salt. Look at the reaction for the dissolving of copper(II) sulfate in water (Equation 4). 

455 www.ckl2.org 



CuSO^s) — » Cu + (aq) +SO4 fa) (Equation 4) 

We begin with copper(II) sulfate as a solid that contains Cu 2+ ions and SO 2 ' ions and it produces Cu 2+ 
ions and S0 2 ~ ions in solution. We dissolved a sample of CuSO^ in a beaker of water. If we look at this 
change, we may first think it is a chemical change because we might think that we have new substances 
being formed. But realistically, if we boiled off the water containing the dissolved copper (II) sulfate ions, 
the CuSO^(s) would remain in the beaker. In other words, we can regain our original starting material by 
a simple physical change, that is, evaporation. Think of the crystal of copper(II) sulfate as a combination 
of the Cu 2+ ions and S0 2 ~ ions that have taken position and then when dissolved in water, lose their 
position. When the water is evaporated, they regain their position. Then the dissolving of an ionic solid 
seems more like a physical change. 

Sample question: Which of the following are physical changes? 

(a) burning paper 

(b) melting wax 

(c) evaporating water 
Solution: 

(a) burning paper is a chemical change (new substances are produced) 

(b) melting wax and (c) evaporating water are physical changes as no new substances are produced, (both 
involve a change in state) 

Solutions of Electrolytes Can Conduct Electricity 

When we started this lesson we said that ionic compounds are held together by electrostatic attraction. 
Arrhenius in the late 1800s first thought that when these ionic compounds dissolved in water, they separated 
from each other into ions and classified them as electrolytes. Since they could conduct electricity in water 
solution, they are considered electrolytes. According to Arrhenius (and current theory - pardon the pun!), 
the ions in solution provided the charged particles needed to conduct electricity. Look at the equation 
below for the dissociation of NaCl. 

NaCl (s) -> Na + (aq) + Cl~ {aq) 

The sodium and chlorine ions are present in the crystal. But once the solid NaCl is added to the water, 
it dissolves, which means that the ions move away from their crystalline structure and are now dispersed 
throughout the water molecules. If two electrodes were to be immersed into a solution of NaCl(aq), 
the Na + {aq) ions would move toward one electrode and the Cl'(aq) ions would move toward the second 
electrode. This movement of the ions allows the electric current to flow through the solution. Therefore, 
NaCl{aq) will behave as an electrolyte and conduct electricity because of the presence of Na + (aq) and 
Cl~{aq) ions. The more ions that are present in solution when the salt dissolves, the stronger the electrolyte 
solution is. 

Sample question: Which of the following will form electrolyte solutions and conduct electricity? 

(a) CaF2 (aq) 

(b) CqHi 2 O g (aq) 

(c) KOH (aq) 

www.ckl2.org 456 



Solution: 

(a) and (c) are solutions that contain positive cations and negative anions that would separate when 
dissolved in water. Since ions are separated in solution, they are electrolytes and will conduct electricity. 

CaF 2 ( s ) ~ * Ca 2+ {aq) + 2 F~(aq) 
Calcium fluoride 
KOH(s) -> K + {aq) + OH~(aq) 
Potassium hydroxide 

(b) Is not an ionic compound but a covalent compound. This means that when it dissolves in water it 
stays together as a molecule and is a non-electrolyte 

C 6 H 12 6 (s) -» C 6 H 12 6 (aq) 

Glucose separate glucose molecules 

Lesson Summary 

• Electrostatic attraction describes the bonding that occurs between the ions of ionic solids. Because of 
the strong electrostatic attraction in ionic solids, ionic compounds tend to have high melting points 
and boiling points. 

• Ionic compounds dissolve in polar solvents, especially water. This occurs when the positive cation 
from the ionic solid is attracted to the negative end of the water molecule (oxygen) and the negative 
anion of the ionic solid is attracted to the positive end of the water molecule (hydrogen) . 

• Electrolyte solutions are ones in which free-moving charged particles can conduct an electrical current. 

Review Questions 

1. Write the reactions for the dissolving of the following. 

(a) NaOH( s) 

(b) LiOH (s) 

(c) C 5 //io0 4 ( iS ) 

(d) NH A Cl {s) 

(e) MgCl 2{s) 

2. Which of the following represent physical changes? Explain. 

(a) explosion of TNT 

(b) dissolving KCI 

(c) sharpening a pencil 

(d) souring milk 

3. Which compound contains electrostatic forces? 

(a) natural gas 

(b) table salt 

(c) air 

(d) sugar 

4. Which of the following is a physical change? 

(a) rotting wood 

457 www.ckl2.org 



(b) rising of bread dough 

(c) rusting iron 

(d) molding cheese 

5. Which of the following is not a physical change? 

(a) melting iron 

(b) pumping gas 

(c) reaction of chlorine with sodium 

(d) reaction of magnesium chloride with water 

6. Which compound is considered to be an electrolyte when dissolved in water? 

(a) HN0 3 

(b) Ci2#22011 

(c) N 2 

(d) CH A 

7. Which compound is not considered to be an electrolyte? 

(a) AgCl 

(b) PbSOi 

(c) C 2 H 6 

(d) HC10 3 

8. Janet is given three solutions. She is to determine if the solutions are electrolytes or not but is not 
told what the solutions are. She makes the following observations. What can you conclude from her 
observations and what help can you offer Janet to determine if the solutions are indeed electrolytes? 



Solution 1 
Solution 2 
Solution 3 



Clear 

Blue but transparent 

Clear 



Further Reading / Supplemental Links 

• http://en.wikipedia.org/wiki 

Vocabulary 

electrostatic attraction When solids form from a metal atom donating an electron (thus forming a 
positive cation) to a non-metal (thus forming a negative anion) the two ions in the solid are held 
together by the attraction of oppositely charged particles. 

chemical changes Changes that occur with the chemical bonding where a new substance is formed. 

physical changes Changes that occur in the physical structure but do not occur at the molecular level. 

electrolyte solutions Solutions that contain ions that are able to conduct electricity. 

16.2 Covalent Compounds in Solution 

Lesson Objectives 

• Understand and describe intermolecular bonds. 
www.ckl2.org 458 



• Understand why molecules stay together when dissolving in solvents. 

• Understand and define non-electrolytes. 

Introduction 

Unlike their ionic counterparts, covalent compounds have a different type of attraction occurring between 
the solute and solvent molecules. As we work through this lesson, think about the periodic table and 
remember that covalent compounds are made up of only nonmetals. This is the foundation for the reasoning 
behind the way covalent compounds dissolve. 

Molecules Held Together by Intermolecular Forces of Attraction 



Unlike ionic bonds that transfer electrons, covalent compounds share electrons to complete an outer electron 
configuration. As a result there are no distinct charges associated with the atoms in covalent compounds. 
Look at the formula for CO2. Carbon has 4 valence electrons and oxygen has 6 valence electrons. 

C : ls 2 2s 2 2p 2 
O : ls 2 2s 2 2p 4 

Carbon will share four electrons with other atoms and oxygen will share two electrons with other atoms. 
In CO2 , each oxygen atom shares two of its electrons with carbon and the carbon shares two of its electrons 
with each oxygen atom. Look at the figure below. 

:o::c:;o: 

• ♦ 

This sharing of valence electrons represents covalent bonding. However, this sharing of electrons may not 
be equal sharing and in the case of carbon and oxygen, it is not equal sharing. An electronegativity table 
lists the electronegativities for elements in the periodic table. Elements with a greater electronegativity 
have a stronger attraction for shared electrons. Therefore they can pull the electrons closer to themselves 
and away from the element that has the more positive electronegativity. For carbon, the electronegativity 
value is 2.5, and for oxygen it is 3.5. 

The reason that oxygen has a higher electronegativity is due to its larger nuclear charge and smaller 
diameter. While both carbon and oxygen have the same shielding of their nuclei, that is, Is 2 , oxygen has 
eight protons in its nucleus while carbon only has six. As a result oxygen will pull any shared electrons closer 
to it. This pulling of electrons is what is being measured in electronegativity. The result in this molecule 
is that the electrons are pulled closer to oxygen than carbon and the resultant structure is represented 
below. 

Polar bonds in a carbon dioxide molecule (but not a polar molecule). (Source: Richard Parsons. CC-BY- 

SA) 






The bonding that occurs within the molecule, as we know, involves intramolecular forces. In this molecule, 

459 www.ckl2.org 



while the sharing of electrons inside the molecule is unequal and therefore polar, the overall result produces 
a non-polar molecule because the shifting of the shared electrons toward the oxygen atoms are in equal 
but opposite directions and there is no dipole moment on the molecule. 

When we dissolve a nonpolar covalent compound, such as CO2, in a non-polar solvent, such as benzene, 
the attraction between the molecules of CO2 and CqHq would be intermolecular forces of attraction. 
The intermolecular attractions that occur between all non-polar substances are London-dispersion forces. 
This similarity of intermolecular forces is what allows these two substances to form a solution. 

Molecules Separate When Dissolved in Solution 

When we studied how ionic solids dissolve, we said that as they dissolve in solution, these solids separate 
into ions. More specifically, ionic solids separate into their positive ions and negative ions in solution. The 
same is not true for molecular compounds. Molecular compounds are held together with covalent bonds 
meaning they share electrons. When they share electrons, their intramolecular bonds do not easily break 
apart, thus the molecules stay together even in solution. For example, when you dissolve a spoonful of 
sugar into a glass of water, the intermolecular bonds are broken but not the intramolecular bonds. 
You can write the following equation for the dissolution of sugar in water. 

Ci2#220n(.y) -» C 12 H220u(a q ) (Equation 1) 

Notice how the molecules of sugar are now separated by water molecules. In other words, sugar's in- 
termolecular bonds are broken but since the molecule has not broken apart, this tells us that the in- 
tramolecular bonds have not broken. Since sugar is a polar compound, its intermolecular bonds are 
dipole-dipole including some hydrogen bonding. The same is true for water and it is this similarity of 
intermolecular bonding that allows the sugar and water to form a solution. Also be sure to take note that 
the sugar did not separate into carbon ions, hydrogen ions and oxygen ions but stayed together as a sugar 
molecule when dissolved in water. This is characteristic of covalent compounds or compounds formed 
between non-metals. 

Sample question: Which compounds will dissolve in solution to separate into ions? 

(a) LiF 

(b) P 2 F 5 

(c) C 2 H 5 OH 
Solution: 

LiF will separate into ions when dissolved in solution: LiF{aq) — » Li + {aq) + F~{aq) 
P2F5 and C2//5O// are both covalent and will stay as molecules in a solution. 

Non-Electrolytes 

Remember that solutions of electrolytes conduct electricity and the conduction is the result of ions moving 
through a solution. With covalent compounds, there are no ions moving around in solution, therefore they 
are classified as non-electrolytes. Non- electrolytes are solutions that do not conduct electricity. If you 
were to connect a conductivity meter to these solutions, there would be no reading or a light would not 
turn on if the wires were placed in a solution containing a non-electrolyte. 

Electrical conduction by an electrolyte. (Source: Richard Parsons. CC-BY-SA) 
www.ckl2.org 460 





Solution of Electrolyte Solution of Non-Electrolyte 

A conductivity apparatus is an incomplete electrical circuit that contains a source of electricity and a light 
bulb or meter that will show when current is flowing through the circuit. The ends of the incomplete 
circuit are prongs that can be lowered into a solution. If the prongs are lowered into a solution containing 
a sufficient number of ions, the circuit will be completed by the solution, current will flow, and the light 
bulb will light up. If the prongs are lowered into a solution with on ions or an insufficient number of ions, 
not enough current will flow to light the bulb. 

Lesson Summary 

• Covalent compounds share electrons to obtain a completed outer energy level. 

• Covalent compounds do not form ions in solution; they stay together as molecules. For example: 

Cl2#220n(. v ) -» Cl2#22<?ll( a? ) 

• Non-electrolytes are solutions that do not conduct electricity. 

Review Questions 

1. Describe the intermolecular bonding that would occur between glucose, C%H\20%, and water. 

2. Define non- electrolyte and give at least one example. 

3. How can you tell by looking at a formula that it is most likely a covalent compound? What does this 
tell you about the bonding? 

4. Describe how you could tell the difference between an electrolyte and a non-electrolyte solution. 

5. Looking at the periodic table, which pair of elements will form a compound that is covalent? 

(a) Ca and Br 

(b) Fe and O 

(c) Si and F 

(d) Co and CI 

6. Which of the following compounds will conduct the least amount of electricity if dissolved in water? 

(a) KN0 3 

(b) BaCl 2 

(c) CsF 

(d) C0 2 

7. Steve is given five solutions in the lab to identify. He performs a conductivity test, a solubility (in 
water) test, crudely measures the hardness of each substance, and determines the melting point using 
a melting point apparatus. Some of the melting points, the teacher tells him are too high or low to 
measure using the laboratory melting point apparatus so she gives him the melting point. For the 
liquids, he determined the boiling points. He gathers all of his data and puts it into a table. His 
teacher gives him the names of the five solutions to match his five unknowns to. Can you help Steve 



461 



www.ckl2.org 



match the properties of the unknowns (from the table below) to the solution names (found under the 
table)? 



Table 16.2: 



Unknown Sub- 


Conductivity 


Solubility (: 


in Hardness 


Melting 


Point 


Boiling 


Point 


stance 




water) 




(°C) 




(°C) 




1 


no (aq) 


soluble 


semi- brittle 


164 








2 


yes (aq) 


soluble 


NA (liquid) 






100 




3 


yes (aq) 


soluble 


brittle 


«800 








4 


no (s) 


insoluble 


soft 


82 








5 


yes (s) 


soluble 


NA (liquid) 






118 





List of Unknown names: 

Sodium chloride 

Naphthalene 

Sucrose 

Hydrochloric acid (dilute) 

Acetic acid 

8. Predict the type of bonding that will form between the elements sulfur and bromine. Will this 
molecule conduct electricity in water solution? 

Further Reading / Supplemental Links 

• http://en.wikipedia.org/wiki 

Vocabulary 

intermolecular bonding The bonding that occurs between molecules. 
non-electrolytes Solutions that do not conduct electricity. 



16.3 Reactions Between Ions in Solutions 

Lesson Objectives 

• Use the solubility chart and/or solubility rules to determine if substances are soluble in water. 

• Use the solubility chart and/or the solubility rules to determine if precipitates will form. 

• Write molecular, ionic, and net ionic equations. 

• Understand the use of ionic and net ionic equations. 

• Identify spectator ions in ionic equations. 



www.ckl2.org 



462 



Introduction 



Many ionic compounds are said to be soluble in water while others are said to be insoluble. However, no 
ionic compound is completely insoluble in water. Every ionic compound dissociates into its ions to some 
extent when placed in water. In fact, the solubility of ionic compounds ranges across a full spectrum from 
as little solubility as 1 X 10 -100 moles/liter to 20 moles/liter. Most chemistry textbooks contain, somewhere 
in them, a chart or table of the solubility of ionic compounds. Most tables divide the ionic compounds into 
two groups, "soluble" or "insoluble". Some books add a third category of "slightly soluble", but this term 
often becomes synonymous with insoluble. It is clear that in order to take a large spectrum of solubilities 
and put them into two categories, a decision must be made on a dividing line. A particular solubility is 
chosen as the dividing line and those compounds whole solubility is below that line are called insoluble and 
those above the line are called soluble. It is possible to find, in different textbooks, solubility tables that 
disagree as to whether or not a particular substance is soluble. If one table had a dividing line of 0.10 M, 
silver acetate would be called insoluble whereas a slightly lower dividing line of 0.075 M would list silver 
acetate as soluble. So, keep in mind that sometimes, it is important to determine the actual numerical 
solubility of a substance instead of to rely on a chart of soluble and insoluble compounds. 

When solutes are dissolved in a solution, the solution is transparent so the dissolved solute particles cannot 
be visually detected. If undissolved particles are present in a liquid, they form a cloudy barrier to light 
passing through the liquid and hence their presence can be detected visually. Eventually, the un-dissolved 
particles will settle to the bottom of the container and then it becomes more apparent that an undissolved 
solid is present. 

When ionic solutions are mixed together, it is often possible to form an insoluble ionic compound even 
though both original compounds were soluble. For example, both silver nitrate and sodium chloride are 
soluble compounds. In 1.0 M solution of these substances, the compounds would be completely dissociated. 
One solution would contain Ag + and NO^. ions and the other solution would contain Na + and Cl~ ions. 
When these two solutions are poured together, all of these ions move around in the solution and come 
into contact with each other. If a sodium ion temporarily attaches to either a chloride ion or a nitrate 
ion, nothing happens because sodium ions are soluble with both these ions and the ions would simply re- 
dissociate. If a silver ion temporarily joins a nitrate ion, nothing happens because silver nitrate is soluble. 
But, when a silver ion comes into contact with a chloride ion, the two ions join together permanently 
because silver chloride is not soluble. When a silver ion combines with a chloride ion, they form an 
insoluble solid particle that will not dissolve. Therefore, when two solutions are mixed, a cloudy, non- 
transparent substance forms and eventually this substance will settle to the bottom of the container. 
When a non-soluble substance is formed in a solution, it is called a precipitate, and the process is called 
precipitation. 

Formation of a precipitate. (Source: Richard Parsons. CC-BY-SA) 



+ 



Solution containing 
Ag and NO^ ions. 




Solution containing 
ft and CrOf" ions. 



Mixture containing K + and NO s in solution 
and an insoluble solid of Ag2CrO- 



463 



www.ckl2.org 



Ions in Solution Can React and Produce Precipitates 

Once a solid substance has been separated into its ions, the ions are then available for reactions. When a 
compound is in the solid state, the ions are held with electrostatic attractions. It must first be dissolved 
in solution to allow ions to move freely. Take for example the reaction between sodium chloride and silver 
nitrate. Both of these compounds are available commonly in the solid form. First, you would take both 
solids and dissolve them in solution, in this case water. (Equation 1 and Equation 2) 



AgNO S ( s) -> Ag+ {aq) + N0 3 {aq) 



(Equation 1) 



NaCl [s) -> Na + (aq) + CI {aq} 



(Equation 2) 



Once the solid dissolves to separate into its ions in solution, these ions are available to react together in 
the chemical reaction (Equation 3). 



AgN0 3(aq) + NaCl [aq) -^ AgCl {s) + NaN0 3{aq) 



(Equation 3) 



Make note that the AgNO^t aq \ and NaCL aq \ reactants show the ions are in solution. In other words 
AgNO?,^) is equivalent to Ag + ( aq j +NC>3~ (aq)-, and NaCl( aq ) is equivalent to Na + ( aq -) +C/~ (aq)- The equation 
represented in Equation 3 is a double displacement reaction which means the cations exchange anions in 
the reactants to form the products. In the laboratory, a precipitate is formed that we determine is silver 
chloride. 

The same reactions can be seen when substances undergo ionization. Remember that ionization forms 
ions in solution. For example, look at Equation 4 for the ionization of sulfuric acid (chemical change) and 
Equation 5 for the dissolving of sodium hydroxide (physical change). Notice how they both end up with 
ions in solution. 



2- 



H 2 S 0^ aq) -> 2 H ( aq) + SO4 (j^ 



NaOH, 



{aq) 



Na- 



irn) 



+ OH 



(aq) 



(Equation 4) 
(Equation 5) 



When sulfuric acid reacts with sodium hydroxide we have a double displacement reaction where the cations 
exchange anions. Equation 6 shows this reaction. 



H 2 S0 4(aq) + 2 NaOH {aq) -* 2 H 2 [L) + Na 2 S0 4(aq) 



(Equation 6) 



Notice in Equation 6, that liquid water is produced, not a solid. As well, the second product is an aqueous 
ionic solution containing the ions Na + ^ and SO^ (aq)- What is essential to these equations and equations 
like these is to visualize the ionic aqueous solutions as ions in solution. 

Most ionic compounds dissociate in water. (Source: Richard Parsons. CC-BY-SA) 



When the text says this 



think this 



I 

AB(aq) 



1 




www.ckl2.org 



464 



Using Solubility Charts and/or Solubility Rules 

How do you know if a precipitate is produced in a double replacement reaction? Not all reactions of this 
type will produce a precipitate. Take for example, if you were to mix a solution of table salt, NaCh s \, and 
Epsom salts, MgSO^, in water, you would not get a precipitate. The reaction is seen below. 

2 NaCl( aq) + MgS 0^ aq) -> Na 2 S A{aq) + MgCl 2 ( aq ) (Equation 7) 

So, how would you know when you would get a precipitate and when not? What scientists use is a set of 
solubility rules or, even better, a solubility chart. Table 16.3 represents the solubility chart for the most 
common cations and anions found in ionic solids. 

Table 16.3: Solubility Chart 





C 2 H 3 0^r- 


col 


cr 


cio- 


CrOf 


I 


NO- 


OH 


o 2 - 


PO\ 


sof 


s 2 - 


Ag+ 


ii 


I 


I 


i 


s 


a 


I 


S 


* 


ii 


I 


ii 


I 


Al 3 + 


S 


s 


* 


s 


s 


* 


S 


s 


/ 


I 


I 


S 


* 


Ba 2 + 


S 


s 


a 


s 


s 


/ 


s 


s 


5 


S 


I 


I 


* 


Ca 2 + 


S 


s 


a 


s 


s 


a 


s 


s 


ii 


ii 


ii 


ii 


ii 


Cu 2 + 


S 


s 


* 


s 


s 


* 


* 


s 


I 


I 


I 


S 


I 


Fe 2+ 


S 


s 


a 


s 


s 


* 


5 


s 


I 


I 


I 


S 


I 


Fe 3+ 


S 


s 


* 


s 


s 


/ 


5 


s 


I 


I 


ii 


ii 


* 


H& 


S 


s 


* 


s 


s 


a 


ii 


s 


I 


ii 


I 


I 


/ 


K + 


S 


s 


5 


s 


s 


s 


s 


s 


S 


S 


S 


S 


5 


Mg 2+ 


S 


s 


a 


s 


s 


a 


S 


s 


I 


I 


ii 


S 


* 


Mn 2 + 


S 


s 


a 


s 


s 


* 


S 


s 


I 


I 


ii 


s 


/ 


Na + 


S 


s 


s 


s 


s 


5 


s 


s 


s 


S 


S 


s 


5 


NH+ 


S 


s 


s 


s 


s 


5 


s 


s 


s 


S 


S 


s 


5 


Pb 2+ 


S 


I 


i 


I 


s 


/ 


I 


s 


ii 


ii 


I 


ii 


/ 


Sn 2+ 


* 


s 


* 


s 


s 


/ 


s 


s 


I 


I 


I 


S 


/ 


Sn i+ 


s 


s 


* 


* 


s 


ii 


* 


s 


ii 


I 


* 


S 


/ 


Sr 2+ 


s 


s 


a 


s 


s 


a 


5 


s 


S 


S 


/ 


ii 


5 


Zn 2 + 


s 


s 


a 


s 


s 


ii 


5 


s 


I 


ii 


/ 


S 


/ 



S = soluble in water, I = insoluble in water, ii = partially soluble in water, * = unknown or does not exist. 

Let's now see how we use the solubility chart to determine if two compounds will form a precipitate 
when they react. If we had a reaction between sodium bromide and silver nitrate, we know that this is 
a reaction between two compounds and therefore is a double replacement reaction. How do we know the 
states of the products formed? The reaction is seen below. 

NaBr {aq) + AgN0 3{aq) -> NaN0 3{?) + AgBr {?) (Equation 8) 

Look at the solubility chart and see if you can predict if the reaction will produce any precipitates. If you 
look across the row for sodium ion, all sodium compounds are soluble (S), therefore you can fill in this 
part of the equation (See Equation 9) 

NaBr {aq) + AgN0 3{aq) -> NaN0 3(aq) + AgBr {?) (Equation 9) 

465 www.ckl2.org 



f 


C,HA' 


Br" 


co a a 


cr 


ao 3 


CrO + 3 


r 


NO/ 


OH" tf- 


?o 4 i 


so 4 2 


S*^ 


*B*— 




I 


I 


i 


s 


ii 


i 


s 


* a 


i 


ii 


I 


V- 




J 



Figure 16.1: Solubility chart being used to determine whether the bromide anion and silver cation combi- 
nation will produce a precipitate. 



If you look across the row for silver, under bromide ion, you find an I for insoluble. Figure 22.4.1 shows 
this row of the solubility chart. 

Therefore we can complete the equation (Equation 10). 

NaBr (aq) + AgN0 3{aq) -> NaN0 3{aq) + AgBr {s) (Equation 10) 

Let's try another one. Take a reaction between ammonium phosphate and lead acetate. 

2 (NH 4 ) 3 PO i{aq) + 3 Pb(C 2 H 3 2 ) 2{aq) -> 6 NH 4 C 2 H 3 2{aq) + Pb 3 (P0 4 ) 2{s) (Equation 11) 

How did we know that the lead(II) phosphate would precipitate from solution? Why did the ammonium 
acetate not precipitate? If we follow the ammonium row in the solubility table across to the acetate column, 
we find an "S" at the intersection which indicates this compound is soluble. If we follow the lead row across 
to the phosphate column, we find an "I" at the intersection, which indicates insolubility and therefore a 
precipitate of this compound will form. 

As we said at the beginning, rather than using a solubility chart, some scientists simply use a set of 
solubility rules. The rules are as follows: 

1. All group 1 metals and ammonium compounds are soluble. 

2. All nitrates, chlorates, and bicarbonates are soluble. 

3. Halides are soluble except for Ag + , Hg 2 + , and Pb 2+ 

4. Sulfates are soluble except for Ag + , Ba 2+ , Ca 2+ , Hg 2 2 + , S r 2+ , and Pb 2+ 

5. Carbonates, chromates, phosphates, and sulfides are insoluble except those from rule #1. 

6. Hydroxides are insoluble except for those in rule #1, and Ba 2+ . 

It is important to remember that this is a priority set of rules. What this means is that Rule #1 is 
first. All group 1 metals and ammonium compounds are always soluble. For example, even though sulfide 
compounds are rarely soluble for any cation (rule #5), they will be soluble with group 1 metal ions or with 
ammonium ions (rule #1). It also does not matter whether you use the set of rules or the solubility chart. 

They both provide the same information; the chart is easier to read for some, the rules are easier to 
remember for others. 

Sample question: Complete the following reactions. Use the solubility table to predict whether precipitates 
will form in each of the reactions. 

(a) Pb(N0 3 ) 2{aq) + KI {aq) -> 

(b) BaCl 2(flq) +Na 2 S0 4(aq) -> 
Solution: 

(a) Pb(N0 3 ) 2{aq) + 2KI (aq) -> Pbl 2(s) + 2KN0 3{aq) 

(b) BaCl 2 ( aq ) +Na 2 SO A ( aq ) -» BaS0 4 ^ + 2NaCl( aq ) 



www.ckl2.org 



466 



Separating Ions in Solution 




When a chemist has a solution that contains two or more ions and he/she wishes to physically separate the 
ions, the differences in solubilities can be used. Suppose that a chemist has a solution that contains both 
Pb 2+ and Zn 2+ ions. If these two ions were dissolved in the solution as nitrates, then the only anion present 
is the nitrate ion. If the chemist added some NaCl to the solution, the zinc ions would remain in solution 
because ZnCl^ is soluble but the lead ions and the chloride ions would form an insoluble compound, PbC^, 
and form a precipitate. The chemist could then pour this mixture through a piece of filter paper and the 
dissolved zinc ions would pass through the filter paper with the solution but the solid PbCli would be 
filtered out. Therefore, the chemist would have separated the zinc ions (now in the solution) and the lead 
ions (now in the filter paper). See Figure ??. 

The process for separating and identifying ions by selective precipitation and filtration is known as gravi- 
metric analysis. Analytical chemistry is a sub-discipline of chemistry. The task of analytical chemists is 
to identify the substances present in materials and to make quantitative measurements on them. In order 
to identify what substances are present, it is often necessary to isolate (separate) the ions from other ions 
that might be present. In earlier times, much of this work was done by gravimetric analysis. In modern 
times, a large amount of the identification of ions in solution is done by instrumentation. 

If you are called upon to determine a process for separating ions from each other, you should look in the 
solubility table to determine a reagent that will form a precipitate with one of the ions but not with the 
other. 

Chemists' knowledge of the interaction of radiation and matter is the basis for analytical methods of 
sensitivity and specificity. Signals from Within (http : //www. learner .org/vod/vod_window.html?pid= 
802) 



467 



www.ckl2.org 



The Reactions Can be Written as Ionic Equations 

Remember earlier when we said that we should visualize NaCh aq \ as Na + r aq \ and Cl~ i aq \ and AgN0 3 ( aq \ as 
Ag + r aq \ and N0 3 ~t aq \? This is so we could write chemical reactions as formula equations. Equation 12 
is a formula equation. 

NaCl {aq) + AgN0 3{aq) -> NaN0 3{aq) + AgCl (s) (Equation 12) 

The ionic solids discussed in the previous section were written together in a formula form. In an ionic 
equation, the separated ions are written in the chemical equation. Let's rewrite Equation 12 as a total 
ionic equation which is a better representation of a double replacement reaction. 

Na + [aq] + Cr (aq) + Ag + {aq) + N0 3 - {aq) -> Na + {aq) + NOf {aq) + AgCl {s) (Equation 13) 

Sample question: Write the ionic equation for each of the following. 

(a) BaCl 2 ( aq ) + Na 2 S Atyaci) -> 2 NaCl^ aq ) + BaS 4(i ) 

(b) 2 K 3 P0 4{aq) + 3 Ca(N0 3 ) 2{aq) -> 6 KN0 3{aq) + Ca 3 (P0 4 ) 2(s) 
Solution: 

(a) Ba 2+ {aq) + 2 CI [aq) + 2 Na + {aq) + S0 4 2 - (aq) -» 2 A^+ K) + 2 C/- (fl?) + BaS0 4{s) 

(b) 6 ^+ H) + 2 POt 3- ^) + 3 Ca 2+ {aq) + 6 MV^) -> 6 K+ {aq) + 6 iV0 3 _ H) + Ca 3 {PO A ) 2(s) 

The Essential Information is Contained in a Net Ionic Equation 

If you look at Equation 13, what do you notice is the same on both sides of the equation? 

Na + {aq) + Cr {aq) + Ag + (aq) + N0 3 ~( aq} -> Na + {aq) + N0 3 ~ \ aq) + AgCl {s) (Equation 13) 

Do you see that Na + r aq \ and N0 3 ~ i aq \ appear on both sides of this equation in the same form? 

©s — ■ v x— v s— n. (Equation 13) 

+ Cl(aq) + Ag(aq) +(No,(aq)]— >(Na (aq))+(No,(aq)J+ AgCl(s) 

These ions, because they appear on both sides of the equation and in the same form, are called spectator 
ions. Think of the phrase "spectator ion". What does it sound like to you? Does it sound like the ion is just 
in the container watching the reaction? Well, that's kind of what a spectator ion is doing. A spectator 
ion is an ion in the ionic equation that appears in the same form on both sides of the equation indicating 
they do not participate in the overall reaction. Therefore Na + (aq) and NO^(aq) are spectator ions for this 
reaction. So what is the overall reaction? Well, let's remove the spectator ions and see. 

A 8 + { aq ) + cr (aq) -» AgCl {s) (Equation 14) 

Without the spectator ions, we see only the ions that are responsible for forming the solid silver chloride. 
This equation, represented in Equation 14, is the net ionic equation. The net ionic equation is the overall 
reaction that results when spectator ions are removed from the ionic equation. The net ionic equation gives 
us all of the essential information we need; what ions we need to form our solid. Really, whether we had 
sodium chloride or potassium chloride is irrelevant, what was important was that the chloride ion and the 
silver ion were present. 

Sample question: Write the net ionic equation for each of the following. Name the spectator ions. 
www.ckl2.org 468 



(a) Ba 2+ {aq) + 2 CI {aq) + 2 Na + (aq) +50 4 2 "( a? ) -» 2 iVa+ (fl9) + 2 C/" W) + flaS0 4( ,) 

(b) 6 ^+ H) + 2 P0 4 3 - K) + 3 Ca 2+ {aq) + 6 tf0 3 'te) -» 6 Z+ (as) + 6 tf0 3 '(a,) + Ca 3 (P0 4 ) 2W 
Solution: 

(a) 

Ba 2+ (aq) + 2 C/ - ^) + 2 iVa + (a(? ) + SO* 2- ^ -> 2 Na + {aq) + 2 C/"( fl9 ) + BaS0 4{s) 

^ fl (a?) + S0 4 ( ac/ ) — » BaSO^ s ) net ionic equation 
Cl~ ( aq ) and Na + ( aq j = spectator ions 

(b) 

6 K + {aq) + 2 P0 4 3 - M + 3 Ca 2+ {aq) + 6 N0 3 ~ {aq) -» 6 tf+ K) + 6 iVOa"^) + Ca 3 (P0 4 ) 2W 

3 Ca 2+ ( aq } + 2 PO^~ t aq \ — > Ca 3 (P0 4 )2(j) net ionic equation 
A" + ( a? ) and NO^~ ( aq ) = spectator ions 

Lesson Summary 

• A solubility chart is a grid showing the possible combinations of cations and anions and their solu- 
bilities in water. It is used to determine whether a precipitate is formed in a chemical reaction. The 
solubility rules are a list of rules dictating which combinations of cations and anions will be soluble 
or insoluble in water. 

• A total ionic equation is one in which all of the ions in a reaction are represented. A net ionic 
equation is one in which only the ions that produce the precipitate are represented. 

Review Questions 

1. What is more valuable to use for determining solubility: a solubility chart or a set of solubility rules? 

2. If you were told to visualize Cu(NOz)2(aq), what might this mean to you? 

3. Use the solubility rules to determine the following solubilities in water. Solubility Rules 

(a) All group 1 metals and ammonium compounds are soluble. 

(b) All nitrates, chlorates, and bicarbonates are soluble. 

(c) Halides are soluble except for Ag + , Hg 2 2+ , and Pb 2+ 

(d) Sulfates are soluble except for Ag + , Ba 2+ , Ca 2+ , Hg 2 2+ , Sr 2+ , and Pb 2+ 

(e) Carbonates, chromates, phosphates, and sulfides are insoluble except those from rule #1. 

(f) Hydroxides are insoluble except for those in rule #1, and Ba 2+ . 

Which of the following compounds is soluble in water? 

(a) PbCl 2 

(b) Hg 2 Cl 2 

(c) (NH A ) 2 SOi 

(d) MgC0 3 

(e) AgN0 3 

(f) MgCl 2 

(g) KOH 
(h) PbSOi 

469 www.ckl2.org 



4. When only the ions that produce a precipitate are shown for a chemical equation, what type of 
reaction exists? 

(a) spectator equation 

(b) molecular equation 

(c) ionic equation 

(d) net ionic equation 

5. If you wanted to separate a solution of Ag 3+ ( aq ^ from a solution of Hg2 2+ \ a q)-, which of the following 
would be the best possible reaction? 

(a) add HBr 

(b) add HN0 3 

(c) add NaOH 

(d) none of the above 

6. If you wanted to separate a solution of Al + ^ from a solution of Fe 2+ ^, which of the following 
would be the best possible reaction? 

(a) add HCl 

(b) add NaOH 

(c) add H 2 S 

(d) none of the above 

7. Complete the following reactions: 

(a) Nci2S(aq) + ZnCh(aq) — > 

(b) (NH^CO^aq) + CaCl 2 (aq) -> 

8. Write the ionic equations for the balanced molecular equations from question 5. 

9. Write the net ionic equations for the ionic equations from question 6. 
10. Identify the spectator ions for the ionic equations from question 6. 

Vocabulary 

solubility chart A grid showing the possible combinations of cations and anions and their solubilities 
in water. 

solubility rules A list of rules dictating which combinations of cations and anions will be soluble or 
insoluble in water. 

formula equation A chemical equation written such that the aqueous solutions are written in formula 
form. 

total ionic equation A chemical equation written such that the actual free ions are shown for each 
species in aqueous form. 

net ionic equation The overall reaction that results when spectator ions are removed from the ionic 
equation. 

spectator ions The ions in the total ionic equation that appear in the same form on both sides of the 
equation indicating they do not participate in the overall reaction. 

www.ckl2.org 470 



Labs and Demonstrations for Ions in Solution 
Teacher's Resource Page for Qualitative Ion Testing Lab 

Investigation and Experimentation Objectives 




In this activity, the student will use self-determined evidence and logically consistent arguments to generate 
conclusions. Detailed record keeping of observations is necessary. 

Lab Notes 

If you have several classes to do the lab, 500 mL of each solution will be adequate. Students have fewer 
spills and create less mess with dropper bottles. You will need to refill dropper bottles from larger bottles 
after each class. If you purchase a large supply of dropper bottles, the labeled bottles can be stored and 
used from year to year. Making and removing labels consumes a large amount of preparation time. 

This lab may require more than one day if your lab periods are 50 minutes or less per day. You can also 
alter the lab so that your students only test for the cations or for the anions but not both. 

Solution Preparation 

3.0 M NH^OH Concentrated NH4OH is 14.5 M so 500 mL can be prepared by diluting 103 mL of concen- 
trated NH4OH to 500 mL. 

3.0 M HNO3 Concentrated nitric acid is 15.6 M so 500 mL can be prepared by diluting 96 mL of concentrated 
HNO3 to 500 mL. 

0.10 M Ba(NOs)2 Dissolve 13.1 grams of Ba{NOs)2 in sufficient water to make 500 mL of solution. Since 
this solution is also used as a testing solution (see below), you can dissolve 26.2 grams in sufficient water 
to make 1.0 L of solution and divide the solution to avoid preparing it again. 

0.10 M KSCN Dissolve 4.86 grams of KSCN in sufficient water to make 500 mL of solution. 

0.10 M AgNOs Dissolve 8.5 grams of AgNOs in sufficient water to make 500 mL of solution. Since this 
solution is also used as a testing solution (see below), you can dissolve 17.0 grams in sufficient water to 
make 1.0 L of solution and divide the solution to avoid preparing it again. 

Unknown salts for students to test, use fiaC/2 or Na2SOi or L/2CO3 (recommended). 

For testing solutions: 

Barium, Ba 2+ , use 0.1 M Ba{NO^)2', Dissolve 13.1 grams of Ba{NO$)2 in sufficient water to make 500 mL 
of solution. 

Iron(III), Fe 3+ , use 0.1 M Fe{NO^)y Dissolve 12.1 grams of anhydrous Fe(7V0 3 ) 3 or 20.7 grams of Fe(NOz)%- 
9H2O in sufficient water to make 500 mL of solution. Lithium, Li + , use 0.1 M LiNOy Dissolve 3.45 grams of 
L1NO3 in sufficient water to make 500 mL of solution. Potassium, K + , use 0.1 M KNO3; Dissolve 5.06 grams 
of KNO3 in sufficient water to make 500 mL of solution. 

Silver, Ag + , use 0.1 M AgNOy, Dissolve 8.5 grams of AgNO^ in sufficient water to make 500 mL of solution. 

Sodium, Na + , use 0.1 M NaNOs; Dissolve 4.25 grams of NaNOz in sufficient water to make 500 mL of 
solution. 

Carbonate, CO^~, use 0.1 M Na2CO%\ Dissolve 5.30 grams of Na2CO% in sufficient water to make 500 mL 
of solution. 

Chloride, Cl~, use 0.1 M NaCl] Dissolve 2.93 grams of NaCl in sufficient water to make 500 mL of solution. 

4T1 www.ckl2.org 



Iodide, / , use 0.1 M Nal; Dissolve 7.50 grams ol Nal in sufficient water to make 500 mL of solution. 

Sulfate, SO 2 ', use 0.1 M Na 2 S0 4 ; Dissolve 7.10 grams of Na 2 SOi in sufficient water to make 500 mL of 
solution. 

For flames tests of Li + , Na + , K + , and Ba 2+ ions, place 25 mL of each of the testing solutions in separate 
150 mL (labeled) beakers and stand 15 wooden splints (one for each lab group) in the beakers to soak. 

Sometimes, the barrel opening of Bunsen burners are contaminated by previous spills and will produce 
colored flames without a testing splint in them. You should check the burners to make sure they do not 
produce colored flames when burning. You should also remind students of burner safety procedures . . . 
no loose hair that falls past your face, etc. 

Answers to Pre-Lab Questions 

1. AgN0 3{aq) + Nal {aq) 
Yes, precipitate forms, Agl 

2. Pb(N0 3 ) 2(aq) + CaCl 2{aq) 
Yes, precipitate forms, PbCl 2 

3. NH A N0 3{aq) + CaCl 2(aq) 
No precipitate forms 

4. Sr{N0 3 ) 2{aq) +K 2 SO i{aq) 
Yes, precipitate forms, SrSO^ 

Qualitative Ion Testing Lab 

Background: 

How are unknown chemicals analyzed? One method is by making comparisons to "known" chemicals. In 
this lab activity, ion tests will be performed and observations made for the reactions of four known anions 
and six known cations. Then an unknown salt will be identified by analyzing and comparing results to 
what is known. 

The process of determining the identities of unknown substance is called qualitative analysis. This can be 
contrasted to quantitative analysis, which is the process of determining how much of a given component 
is present in a sample. Qualitative analysis procedures use physical tests as well as chemical tests. The 
physical tests in this lab involve observing colors of solutions and colors produced in flame tests. The 
chemical tests in this lab involve chemical reactions, as evidenced by formation of a precipitate, dissolving 
of a precipitate to form a complex ion, a color change, or evolution of a gas. 

Formation of a Precipitate 

An ionic salt is a compound composed of two parts - cations (positively charged ions) and anions (negatively 
charged ions). When an ionic salt is dissolved in water, the salt crystal dissociates or separates into its 
cations and anions. For example, potassium iodide (KI) dissociates into potassium ions (K + ) and iodide 
ions (/"") according to equation 1. 

KI (*) -» K tad) + 7 K) Equation 1 

Similarly, the ionic salt lead(II) nitrate, Pb(N0 3 ) 2 , dissociates into lead cations, Pb 2+ , and nitrate anions, 
N0 3 , according to equation 2. 

Pb(N0 3 ) 2(s) -» Pb 2 + } + 2 NO- (aq) Equation 2 

www.ckl2.org 472 



When two ionic salts are mixed together in water, two new combinations of cations and anions are possible. 
In some cases, the cation from one salt and the anion from the other salt may combine to form an insoluble 
product, which is called a precipitate. For example, if solutions of potassium iodide and lead(II) nitrate 
are mixed together, a solid precipitate of lead(II) iodide, Pbl2, forms, as shown in equation 3. 

2 K U + 2 7 H) + Pb H) + 2 NO ka t ) - p bh(s) + 2 K+ q) + 2 N0 3{aq) Equation 3 

Notice that the potassium cations and the nitrate anions remain dissolved in solution. They did not change 
during the reaction and are therefore, referred to as spectator ions. In net ionic equations, spectator ions 
are omitted. A net ionic equation is one that includes only the ions participating in the reaction. Thus, 
equation 3 can be reduced to equation 4. 

Pb \a q ) + 2 7 M "» PM ^) EqUatIOn 4 

Dissolving Precipitates through Complex-Ion Formation 

A complex ion is a water-soluble, charged species containing a central atom and other molecules bonded to 
it. The formation of a complex ion is commonly evidenced by the dissolution of a precipitate. For example, 
copper (II) hydroxide, Cu(OH) 2 , is insoluble in water but will dissolve when excess ammonia is added to 
it, forming a soluble copper amine complex ion, Cu(NH 3 ) 4 + , according to equation 5. 

Cu(OH) 2{s) + 4 NH 3{aq) -» Cu(NH 3 f 4 +{aq) + 2 OH~ {aq) Equation 5 

Evolution of a Gas 

Certain anions, such as the carbonate ion, CO 2- , and sulfide ion, S 2 ~ , evolve gas when treated with a dilute 
strong acid. For example, the reaction of calcium carbonate, CaC0 3 , with nitric acid, HN0 3 , produces 
carbon dioxide gas, C0 2 , according to equation 6. 

CaC0 3(s) + 2 H+ -> Ca 2 + + C0 2{g) + H 2 (L) Equation 6 



Flame Colors 



Some metallic salts will display a distinctive color of light when placed in a flame. When the colored light 
from any one of these flames is passed through a prism or viewed through a diffraction grating, a portion 
of the spectrum is visible, containing only a few colors at specific wavelengths, including the colors in the 
original flame. A partial spectrum that contains only discrete lines is called a line spectrum. When heated 
in a flame, electrons in the metal absorb energy from the flame and are promoted to excited energy levels. 
They emit light as they relax back down to the ground state. Each line in the spectrum represents a 
different electronic transition. Since each element has a unique electronic configuration, an element's line 
spectrum, and thus its flame color, is unique and can be used for identification. 

Pre-Lab Questions 

On the last page of this laboratory packet is a listing of solubility rules. Use those rules to determine which 
of the following mixtures of solutions would produce a precipitate. Write the formula for the precipitate 
where one forms. 

1. AgN0 3{aq) +NaI {aq) 

2. Pb{N0 3 ) 2{aq) + CaCl 2{aq) 

3. NH 4 N0 3{aq) + CaCl 2(aq] 

4. Sr{N0 3 ) 2{aq) +K 2 SO i{aq) 

4T3 www.cki2.0rg 



Purpose 

In parts I and II of this lab, qualitative tests for four known anions and six known cations will be performed. 
Test results will be noted and recorded. In part III, the same tests will be performed on an unknown ionic 
salt which contains one of the six possible cations and one of the four possible anions. The cation and 
anion that make up the unknown salt will then be identified. 

Apparatus and Materials 

All solutions should be in drop control dispenser bottles. 

Ammonium hydroxide, NH4OH, 3.0 M 

Barium nitrate solution, Ba{NOs)2, 0.10 M 

Nitric acid solution, HNO3, 3.0 M 

Potassium thiocyanate solution, KSCN, 0.10 M 

Silver nitrate solution, AgNOj, 0.10 M 

Unknown salt solution 

Distilled water 

Beral-type pipets, labeled 

Bunsen burner setup 

Reaction plate, 24-well 

Sheet of notebook paper 

Wooden splints 

Beakers, 250 - mL, 15 each 

Watchglass, 15 each 

Cation Testing Solutions (0.10 M solutions of the nitrates) 

• Barium, Ba 2+ 

. Iron (III), Fe i+ 

• Lithium, Li + 

• Potassium, K + 

• Silver, Ag + 

• Sodium, Na + 

Anion Testing Solutions (0.10 M solutions of the sodium or potassium compounds) 

• Carbonate, C0 2 ~ 

• Chloride, CI 

• Iodide, I~ 

. Sulfate, SOf 

Safety Issues 

All solutions are irritating to skin, eyes, and mucous membranes, particularly the 3.0 M NH4OH and HNO3 
solutions. The silver nitrate solution will turn skin and clothes permanently black. Handle solutions with 
care, avoid getting the material on you, and wash your hands carefully before leaving the lab. As with all 
labs, do not mix any chemicals other than the ones you are directed to mix. 

Procedure 

Part I - Anion Testing for Cl~, I , SO 2 ^ and CO\r 

Preparing the well plate: 

www.ckl2.org 4T4 



1. Obtain a 24- well reaction plate and set it on the lab bench with a piece of white paper underneath it. 
Label the paper as shown in Data Table 1. Notice that the 24- well reaction plate is divided into 6 columns 
(1-6) and 4 rows (A-D). 

2. Using a pipet, add 5 drops of the Cl~ anion-testing solution to wells A, B, and C of column 1. 

3. Using a pipet, add 5 drops of the I~ anion-testing solution to wells A, B, and C of column 2. 

4. Using a pipet, add 5 drops of the SO|~ anion-testing solution to wells A, B, and C of column 3. 

5. Using a pipet, add 5 drops of the CO 2 ' anion-testing solution to wells A, B, and C of column 4. 
Performing the tests: 

Silver Nitrate Test 

6. Add 3 drops of 0.10 M AgNO^ to the first four wells across row A. Observe the formation of precipitates 
and/or color changes. You may need to remove the paper to see clearly. Record detailed observations in 
Data Table 1. 

7. Add 5 drops of 3.0 M HNO3 to each of the precipitates from step 6. Gently swirl the well plate to stir. 
Observe which precipitates dissolve and which do not. Record observations in Data Table 1. 

8. Add 10-12 drops of 3.0 M NH^OH to each of the remaining precipitates from step 6. Gently swirl and 
observe which precipitates dissolve and which do not. Record observations in Data Table 1. 

Barium Nitrate Test 

9. Add 3 drops of 0.10 M Ba{NO^)2 to the first four wells across row B. Observe the formation of precipitates 
and/or color changes. You may need to remove the paper to see clearly. Record detailed observations in 
Data Table 1. 

10. Add 5 drops of 3.0 M HNO3 to each of the precipitates from step 9. Gently swirl the well plate to stir. 
Observe which precipitates dissolve and which do not. Record observations in Data Table 1. 

11. Add 3 drops of 3.0 M HNO3 to to the first four wells across row C. Make observations, looking for the 
strong evolution of gas bubbles. Record observations in Data Table 1. 

12. Repeat any tests for which results were unclear in Row D of the well plate. Rinse the plate with plenty 
of tap water and then rinse with deionized water to prepare the plate for part II. 

Part II - Cation Testing for Li + , Na + , K+, Ag + , Ba 2+ , and Fe 3+ . 

Preparing the Well Plate 

13. Obtain a 24-well reaction plate and set it on the table with a piece of notebook paper underneath it. 
Label the paper as shown in Data Table 2. 

14. Using a pipet, add 5 drops of the Li + cation-testing solution to the top well (A) of column 1. 

15. Using a pipet, add 5 drops of the Na + cation-testing solution to the top well (A) of column 2 

16. Using a pipet, add 5 drops of the K + cation-testing solution to the top well (A) of column 3. 

17. Using a pipet, add 5 drops of the Ag + cation-testing solution to the top well (A) of column 4. 

18. Using a pipet, add 5 drops of the Ba 2+ cation-testing solution to the top well (A) of column 5. 

19. Using a pipet, add 5 drops of the Fe 3+ cation-testing solution to the top well (A) of column 6. 
Performing the Tests: 

20. Observe each solution and record the color of each solution in Data Table 2 (Row A). 
Potassium Thiocyanate Test 

21. Add 3 drops of 0.10 M KSCN to each of the six wells across row A. Gently swirl the plate to stir. 

475 www.ckl2.org 



Observe the formation of precipitates and/or color changes. Record detailed observations in Data Table 2. 
(Row B) 

Flame Tests 

Note: Several of the cations may be identified using flame tests. The flame tests will be performed on the 
four cations that did not show a reaction in step 21. 

22. Set up a Bunsen burner. Adjust the air so the flame color is blue (NO YELLOW) and a distinct inner 
blue cone is apparent. 

23. Obtain a wooden splint which has been soaking in the Li + cation testing solution for at least 15 minutes. 

24. Hold the wooden splint in the flame, flat side down. (Do not touch the top of the burner with the 
wooden splint.) The top end of the splint should be placed directly into the inner blue cone. A distinct 
color should be apparent. Record the flame color in Data Table 2 (Row C). Do not hold the splint in the 
flame too long or the splint will begin to burn. 

25. Repeat steps 24 and 25 using the Na + cation-testing solution, then the K + cation-testing solution, and 
finally the Ba 2+ cation-testing solution. Be careful not to touch the splints together when gathering them. 

26. Rinse the well-plate in the sink with plenty of tap water and make a final rinse with deionized water. 
The splint should be discarded in the waste basket. 

Part III — Identification of an Unknown Salt 

Note: the unknown salt is made up of one of the cations and one of the anions previously tested. 

27. Obtain from your teacher a pipet filled with an unknown salt solution and a pre-soaked wooden splint 
of the same unknown salt. Be sure to record the unknown identifying letter in the Data Tables. 

28. Determine the identity of the cation and the anion that make up the unknown salt. To do this, repeat 
the steps in Part I and Part II. Record all observations for anion testing in Data Table 1 and for cation 
testing in Data Table 2. 

Data for Anion Testing 

Record detailed observations inside the circles on the table. Record all colors that form. Record whether 
any gases evolve. If any solid precipitates form, use the abbreviation PPT. If no reaction occurs, use the 
abbreviation NR. 



www.ckl2.org 476 



Crankm I anion SO 4 anion CO 3 anion Unknown _ Unknown_ 

12 3 4 5 6 

-000000 

■000000 
000000 

000000 



Data for Cation Testing 



In row A, put the original color of the solution. 

In row B, reaction with KSCN. 

In row C, flame test color. 

Row D is for unknown testing. In column 1, put the original color of the solution. In column 2, 

put the results of reaction with KSCN. In column 3, put the flame test color. If you do a second 

unknown, use columns 4, 5, and 6. 

477 www.ckl2.org 



CI" anion I" anion S0 4 2 " anion COg 2 " anion Unknown Unknown 

12 3 4 5 6 

oooooo 
■oooooo 
oooooo 
oooooo 

Post-Lab Questions 

1. Write the net ionic equation for each precipitation reaction that occurred in Parts I and II. Include the 
well identification (Al) as shown in the example below. 

Part I - Anion testing 

Well Al Ag + + CI -> AgCl( s ) (white ppt) 

2. Unknown letter 

(a) What cation is present in your unknown? 

(b) What anion is present in your unknown? 

3. Write the name and formula for your unknown salt. 
Solubility Rules 



• 1. All group 1 metals and ammonium compounds are soluble. 

• 2. All nitrates, chlorates, and bicarbonates are soluble. 

• 3. Halides are soluble except for Ag + , Hg 2 + , and Pb 2+ . 

• 4. Sulfates are soluble except for Ag + , Ba 2+ , Ca 2+ , Hgl + , Sr 2+ , and Pb 2+ . 

• 5. Carbonates, chromates, phosphates, and sulfides are insoluble except those from rule #1. 

• 6. Hydroxides are insoluble except for those in rule #1, and Ba 2+ . 

www.ckl2.org 478 



Conductivity of Solutions Demonstration 

Brief description of demonstration 




Electrodes of a bulb-type conductivity tester are submerged into acid solutions of increasing strength but 
equal concentration. The bulb glows brighter with the increasing acid strength. 

Materials 



100 mL distilled water 

100 mL 0.1 M hydrochloric acid, HCI 

100 mL 0.1 M acetic acid, HC 2 H 3 2 

100 mL 0.1 M citric acid, H 3 C 6 H 5 7 

100 mL 0.1 M malonic acid, H 2 C 3 H 2 04 

100 mL 0.1 M ascorbic acid, HC e H 7 O e 

100 mL 0.1 propanoic acid, HC3H5O2 

100 mL 0.1 glycine, HC 2 H 4 2 N 

100 mL alanine, HC 3 H 6 2 N 

2 mL universal indicator solution, 1-10 pH range 

9 250 mL beakers 

disposable pipette 

8 stirring rods 

conductivity tester, light bulb type 

wash bottle 

400 mL beaker 



Procedure 

Label each of the beakers with the appropriate acid names (or distilled water). Pour 100 mL of each acid 
into separate beakers. Add 4 drops of universal indicator solution to each beaker and stir. Arrange the 



479 



www.cki2.0rg 



beakers in order of the spectrum, from red to yellow. A white background behind the beakers will help color 
definition. Place the electrodes of the conductivity tester into the glycine solution. Plug the conductivity 
tester in. The bulb will glow dimly. (Darken the room if necessary to see it.) Unplug the conductivity 
tester, and rinse the electrodes with a wash bottle into a 400 mL beaker. Repeat this procedure with 
the other acid solutions, noting that the conductivity and thus the brightness of the bulb increases with 
increasing acid strength. 

Hazards 

All of the acid solutions are corrosive, HCl especially so because it is a strong acid. Avoid contact. The 
conductivity tester is a considerable electrical shock hazard, especially with the solutions of electrolytes. 
Make sure that it is unplugged before handling it. 

Disposal 

Rinse each acid solution down the sink with a 100 fold excess of water. 



Image Sources 

(i) • 



www.ckl2.org 480 



Chapter 17 



Acids and Bases 



17.1 Arrhenius Acids 

Lesson Objectives 

• Define an Arrhenius acid and know some examples of acids. 

• Define operational and conceptual definition. 

• Explain the difference between operational and conceptual definitions. 

• Describe the properties of acids. 

• Describe some of the reactions that acids undergo. 

Introduction 

In previous chapters you may remember learning about electrolytes and you may even remember reading 
about Svante Arrhenius. Arrhenius worked a great deal with specific types of electrolyte solutions known 
as acids and bases. He set the groundwork for our current understanding on acid-base theory. We will 
begin our study of acids and bases with Arrhenius's theories starting with his famous definitions. This was 
quite an accomplishment for a scientist in the late 19' ft century with very little technology but with the 
combination of knowledge and intellect available at the time. Arrhenius led the way to our understanding 
of how acids and bases differed, their properties, and their reactions. 

We may not realize how much acids and bases affect our lives. Have you ever thought of drinking a can 
of soda pop and actually drinking acid? Have you looked at bottles of household cleaners and noticed 
what the main ingredients were? Have you ever heard a shampoo commercial and heard them say that the 
shampoo was "pH balanced" and wondered what this meant and why it is so important for hair? Thanks 
to the beginning work of Arrhenius in the latter part of the 19 century, we started to learn about acids 
and bases; our study continued and is constantly growing. Let's begin our study of this wonderful branch 
of chemistry. 

Definition of Arrhenius Acid 

Take a look at all of the following chemical equations. What do you notice about them? What is common 
for each of the equations below? 

481 www.ckl2.org 



HCl(aq) -> H+ {a q ) + CI {aq) (Equation 1) 

Hydrochloric acid 

HN0 3(aq ) -> H + (aq) + N0 3 - {aq) (Equation 2) 

Nitric acid 

HCl0^ aq) -> H + {aq) + ClO^(a q) (Equation 3) 

Perchloric acid 

One of the distinguishable features about acids is the fact that acids donate H + ions in solution. If you 
notice in all of the above chemical equations, all of the compounds ionized to produce H + ions. This is 
the one main, distinguishable characteristic of acids and the basis for the Arrhenius definition of acids. 
Arrhenius defined an acid as a substance that produces H + ions in solution. 

Most acids can be easily identified because their formula begins with H. Notice this is the case with the 
three acids in the equations above. This, of course, is not always the case. Sometimes acids are written 
a little differently so that the H is not written first. Let's look at the chemical equation for acetic acid 
(Equation 4). It can sometimes be written in a different manner (Equation 5), which is typical for weak 
acids. 

HC 2 H 3 2 (aq) -> H+ {a q ) + C 2 H 3 2 ~ ( aq ) (Equation 4) 

CH 3 COOH (aq) -» H+ (aq) + CH 3 COO- {aq) (Equation 5) 

Thus, there are two ways you could identify an acid. You could check to see if the compound formula 
begins with an H, this would be a primary indicator. Then write the ionization equation, if the ionization 
equation reveals that the H + ion is released, the compound is definitely an acid. 

Sample question 1: Which of the following compounds are acids? For those that are acids, write the 
ionization reaction. 

(a) H 2 SO A 

(b) NaOH 

(c) C 6 H 5 COOH 

Solution: 

(a) H 2 S 04 looks like it is an acid because the formula begins with an H. We check by writing the ionization 
equation and see that the compound dissociates to give H + ions and therefore is definitely an acid. 

H 2 S0 4 ( ag ) -» 2 H ( aq) + S0 4 "(^ 

(b)NaOH has Na + as a cation, not H + (or starts with a cation other than H + ) and is therefore not an 
acid. By writing the dissociation equation we see that NaOH is definitely not an acid. 

NaOH {s) -> Na + {aq) + OH~ {aq) 

(c) CqH^COOH does not start with hydrogen but when we write the ionization equation, we reveal that 
the compound ionizes to give H + ions in solution and is therefore an acid. 

C 6 H 5 COOH( aq j t? H+ (aq) + CqH 5 COO~ ( aq ) 
www.ckl2.org 482 



Properties of Acids 

Acids are a special group of compounds because it has been found that they have their own set of properties. 
This helps to identify them from other compounds. Thus, if you had a number of compounds that you 
were wondering whether these were acids or otherwise, you could identify them by their properties. But 
what exactly are the properties? Think about the last time you tasted lemons. Did they taste sour, sweet, 
or bitter? Lemons taste sour. This is another property of acids. Another property of acids is that they 
turn blue litmus paper red. Litmus paper is an indicator paper that is used to identify whether a substance 
is an acid or a base. If blue litmus paper turns red when it is dipped into a solution, then the solution is 
an acid. Figure ?? shows litmus paper and its reaction with an acid solution. 




The properties of acids such as sour taste and turning blue litmus red are parts of the operational definition 
of acids. Operational definitions describe how the acids behave. Operational definitions differ from 
conceptual definitions because with conceptual definitions, there is an attempt to explain why the acid 
is behaving the way it is. Let's look at the reaction between hydrochloric acid and sodium hydroxide. 
Hydrochloric acid ionizes and produces an H + ion. 



HCl 



(aq) 



/r 



i a i) 



+ cr 



(aq) 



(Equation 1) 



Hydrochloric acid 



Sodium hydroxide, on the other hand, dissociates to produce the sodium ion and the hydroxide ion. Look 
at Equation 6. 

NaOH^ -> Na + ^ + OH~^ (Equation 6) 

When hydrochloric acid and sodium hydroxide react, we get the following chemical equation (Equation 7). 

HCh aq \ + NaOH( aq -j — > NaCl( aq j + //20(l) (Equation 7) 

Notice how the products of Equation 7 are table salt and water. When an acid (HCl) reacts with a base 
(NaOH), a salt and water are produced. Therefore the acid has neutralized the base. The conceptual 
definition for the neutralization would state that the H + ions and the OH ions react to form neutral 
water molecules. This is true for all reactions between acids and bases. 

Sample question: Write the reaction between HNO3 and KOH. Explain how the acid is neutralizing the 
base. 

Solution: 



HN0 3{aq) + KOH {aq) -* KN0 3{aq) + HOH {L) 

483 



www.cki2.0rg 



As with the case with HCl and NaOH, the H + ions from HN0 3 r aq \ and the OH~ r aq \ ions from the KOHi aq \ 
combine to form neutral water. This means that the acid has neutralized the base. If we were to write the 
total ionic equation for this reaction, we would see: 

Total Ionic: H+ {aq) + N0 3 ~ (aq) + K+ {aq) + OH~ {aq) -> K + {aq) + N0 3 " {aq) + H 2 {L) 

The net ionic equation then reveals that the acid neutralizes the base. Net ionic: H + t aq \ + OH~ i aq \ — » H 2 On\ 

Acids React with Metals 

We have learned that part of the conceptual definition of acids is that they neutralize bases. Another part 
of this definition is that they react with metals to produce hydrogen gas. Look at the chemical reactions 
below. What type of reactions are these? 

Z« w + 2 HClfa) -> ZnCl 2 ( aq ) + H 2 ( g ) (Equation 8) 

Mg {s) + 2 HCl {aq) -> MgCl 2{aq) + H 2{g) (Equation 9) 

Ba( s ) + 2 HCl( aq j — > BaClzUq) + #2(g) (Equation 10) 

Notice that each of these reactions is a single replacement reaction. Single replacement reactions involve 
the reaction between a single element and a compound. In the cases shown above, the single displacement 
reactions all involve a metal reacting with an acid. What do you notice that is the same in the product 
side of the equation for all three equations? They all produce hydrogen gas (H 2 ). This is another part of 
the conceptual definition of acids. Acids react with most metals to produce hydrogen gas. 

Sample question: Write the reaction between: 

(a) magnesium and sulfuric acid. 

(b) calcium and acetic acid 
Solution: 

(a) Mg (s) + H 2 SO A{aq) -> MgSO A{aq) + H 2(g) 

(b) 2 CH 3 COOH {aq) + Ca {s) -» Ca(OOCCH 3 ) 2{aq) + H 2{g) 

Acids in Our Environment 

Acids are present in our everyday lives. Think about the last time you took an aspirin or a vitamin C 
tablet. Aspirin is acetylsalicylic acid while vitamin C is ascorbic acid; both are acids that can produce H + 
ions when ionizing in water. Acetic acid {HC 2 H 3 2 ) is a component of vinegar, hydrochloric acid {HCl) is 

stomach acid, phosphoric acid (//3PO4) is commonly found in dark soda pop, sulfuric acid (H2SO4) is used 
in car batteries and formic acid {HC0 2 H) is what causes the sting in ant bites. The list goes on and on. 
We interact with acids on a daily basis so some knowledge of their properties and interactions is essential. 

Sample question: Write the names of three common acids other than the ones listed here. 

Solution: 

Hydrofluoric acid, HF - used to etch glass 

Nitric acid, HNO3, - used to etch metals 

Citric acid, C 3 H§{COOH) 3 - sour taste in citrus fruits 

www.ckl2.org 484 



Lesson Summary 

• Arrhenius defined an acid as a substance that donates H + ions when dissociating in solution. Oper- 
ational definitions describe observed properties of how acids behave. 

• Conceptual definitions describe why acids behave the way they do. 

• Properties of acids include: affect indicators (turn blue litmus paper red), taste sour, neutralize bases, 
react with metals to produce hydrogen gases. 

Review Questions 

1. Explain the difference between a conceptual definition and an operational definition. 

2. What are the properties of acids? Give a common example. 

3. Which statement best describes a characteristic of acid solutions? 

(a) They react with some metals to form hydrogen gas. 

(b) They turn red litmus paper blue. 

(c) They taste bitter. 

(d) They are made from non-metal oxides. 

4. Which of the following is the Arrhenius definition of an acid? 

(a) An acid is a substance that donates protons. 

(b) An acid is a substance that accepts protons. 

(c) An acid is a substance that dissolves in water to form OH ions. 

(d) An acid is a substance that reacts with water to form H + ions. 

5. Which of the following will react with acids and produce hydrogen gas? 

(a) chlorine 

(b) ammonia 

(c) carbon 

(d) magnesium 

6. Write the reaction for each of the following: 

(a) hydrofluoric acid + sodium hydroxide 

(b) potassium hydroxide + hydrogen sulfide 

(c) dissociation of iodic acid 

(d) zinc + hydrochloric acid 

Further Reading / Supplemental Links 

• http://en.wikipedia.org/wiki 

Vocabulary 

Arrhenius acid A substance that produces H + ions in solution. 

operational definitions Definitions that describe how something behaves, (i.e. the operational defini- 
tion of acids includes tastes sour and turns blue litmus red). 

conceptual definitions Definitions that describe why something behaves the way it does. (i.e. the 
conceptual definition of acids includes reacting with bases to neutralize them) . 

485 www.ckl2.org 



17.2 Strong and Weak Acids 

Lesson Objectives 

• Distinguish between strong and weak acids. 

• Identify strong and weak acids from given choices. 

• Describe how strong and weak acids differ in terms of concentrations of electrolytes. 

Introduction 

A great number of people associate a strong acid with its ability to react with skin, essentially "melting ' 
it away from bone. It was only recently on a popular crime show that this very acid chemistry know-how 
was used as a method for a crime. This crime show used sulfuric acid. Why sulfuric acid and not acetic 
acid? What makes the difference? How can we tell if an acid is strong or weak? How does this relates to 
the electrolyte lesson we have learned previously, considering that the acids are a combination of the H + 
cation and the anion? The answers to these questions and more will be found in the lesson that follows. 

The Hydronium Ion 

As has been discussed in an earlier chapter, ions in solution are hydrated (Figure 17.1). That is, water 
molecules are attached to the ions by the attraction between the charge on the ion and oppositely charged 
end of the polar water molecules. A positive ion in solution will be surrounded by water molecules with 
the negative ends of the polar water molecule oriented toward the positive ion. A negative ion in solution 
will be surrounded by water molecules with the positive ends of the water molecules oriented toward the 
negative ion. When we write the formula for these ions in solution, we do not show the attached water 
molecules. It is simply recognized by chemists that ions in solution are always hydrated. 





Figure 17.1: Hydrated ions in solution. 

The hydrogen ion in solution is also hydrated in the same as all other ions. Many chemists feel that it 
is important in the case of hydrogen ion to show that it is attached to water. They choose not to show 
that the hydrogen ion is surrounded by four water molecules, but show only that it has one water molecule 
attached. When you add a hydrogen ion, H + to one water molecule, H2O, the result is a positively charged 
ion consisting of three hydrogens and one oxygen, H^O + . This ion has been given the name hydronium 
ion. Many authors show this hydronium ion in every case where a hydrogen ion is represented in an 
equation, but other authors show it only sometimes. Expressing the hydrogen ion as a hydronium ion 

www.ckl2.org 486 



complicates equations somewhat because if the hydrogen ion is shown as a hydronium ion, then the other 
side of the equation must contain a water molecule to balance the equation. 



HCl 



(aq) 



/T 



(aq) 



+ cr 



(aq) 



(not showing hydronium) 



HCl(aq) + #20(L) -> H 3 ( ag ) + CI ( aq ) 



(showing hydronium) 



You will need to recognize that sometimes the hydrogen ion is shown simply as a hydrogen ion and 
sometimes, it is shown as a hydronium ion and you will need to be able to deal with it in either form. 

Strong Acids 

In an earlier chapter, ionization was defined as certain covalently bonded molecules reacting with water and 
producing ions in solution. Ionization reactions involve chemical changes. It is the amount of ionization 
that is essential in determining if an acid is strong or weak. Strong acids are defined are ones that 
completely ionize or undergo 100% ionization (See Equation 1). 



HCl (g) + H "zO(L) -* H 3 (aq) + CI ( aq ) 



(Equation 1) 



Notice in equation 1 that there is a single arrow separating the products from reactants. This single arrow 
indicates that when the reaction has stopped, there are no HCh g \ molecules remaining in the solution, only 
H^ (aq) i° ns an d Cl~f aq \ ions. This is characteristic of a strong acid. 

There are only six common strong acids. These acids are shown in Table 17.1. Each of the acids found in 
this table, like HCl, completely ionize in water. 

Table 17.1: Strong Acids 



Name 



Symbol 



Hydrochloric Acid 
Hydrobromic Acid 
Hydroiodic Acid 
Nitric Acid 
Perchloric Acid 
Sulfuric Acid 



HCl 

HBr 

HI 

HN0 3 

HClOi 

H 2 SOi 



Weak Acids 



Unlike their strong acid counterparts, weak acids do not ionize 100%. The less ionization that takes place, 
the weaker the acid since there will be fewer H + ions in solution. For example, acetic acid ionizes only 
about 5% meaning that when acetic acid is placed in water, only about 5% of the acetic acid molecules 
separate into H + ions and C2//3O2 ions (See Equation 2). 



HC2H 3 2 ( aq ) ^=> H (aq) + C2H3O2 (aq) 



(Equation 2) 



Notice, as well, that in the above equation our arrow has been replaced with a double arrow indicating 
that the reaction reaches equilibrium. When this reaction reaches equilibrium, the container holds mostly 
acetic acid molecules and a little bit of hydronium (or H + ) ions and a little bit of acetate ions. 



487 



www.ckl2.org 



Let's look at citric acid. Citric acid, C^figO-?, is commonly found in everyday products like lemons and 
limes, and even soft drinks. It is the substance responsible for making the sour taste of these foods and 
drinks. Citric acid only ionizes a little more than 3% and is therefore classified as a weak acid. If we were 
to write an ionization reaction for citric acid, it would appear as written in Equation 3. 

C 6 Hg0 7 (aq) + H 2 O tyVj i? H 3 + (a? ) + CqH 7 7 ~ \ aq ) (Equation 3) 

Notice how, as with Equation 2, the double arrow indicates the reaction has reaches equilibrium. As with 
all weak acids, when this reaction reaches equilibrium, there is mostly C&HgO-jt aq \ and some hydronium 
ion and some citrate ion, C%HjOf~ ' r aq \ in the solution. 

Remember that all acids that are not one of the six listed in Table 1 are weak. These weak acids do not 
completely ionize in water. Even though these weak acids are very soluble, they dissolve as molecules and 
only a few of the molecules break into ions in the solution. 

Sample question: Write ionization equations for only those acids that are weak. 

(a) Sulfuric acid (H2SO4) 

(b) Hydrofluoric acid (HF) 

(c) Trichloroacetic acid (CCI3COOH) 
Solution: 

(a) H2SO4 is a strong acid (one of the six). 

(b) HF is a weak acid (not one of the six): HF{aq) t> H + (aq) + F~{aq) 

(c) CCI3COOH is a weak acid (not one of the six): CCl^COOH^ ±5 H + (aq) + CCl^COO~ 7^) 

Strong and Weak Electrolytes 

Non-electrolytes have been described as solutions that do not conduct electricity and electrolytes are those 
that do conduct electricity. However, electrolytes do have varying degrees of strength. If a solution has a 
large number of ions present in it, it is called a strong electrolyte whereas an electrolyte solution that has 
only a few ions present is called a weak electrolyte. A strong acid completely ionizes in water solution, 
there are lots of ions in solution and the ions make the solution a strong electrolyte. A weak acid produces 
only a few ions in solution and therefore is classified as a weak electrolyte. Such a solution will only conduct 
a small electric current. 

Lesson Summary 

• Strong acids undergo 100% ionization in water (i.e. hydrochloric acid, HCL). Weak acids undergo 
less than 100% ionization (i.e. acetic acid, //C2//3O2). 

• All acids that are not one of the six are weak and also if they are not one of the six, do not completely 
ionize in water. *Weak electrolytes are solutions that conduct electricity to a lesser extent than strong 
electrolyte solutions. 

Review Questions 

1. What is the difference between a strong and weak acid? Show an example of each. 

2. In terms of electrolyte solutions, how would you distinguish between a strong acid and a weak acid? 

3. All of the following are weak acids except? 

www.ckl2.org 488 



(a) HCIO3 

(b) HC 2 H 3 2 

(c) HF 

(d) //C/ 

4. Which compound is a strong acid? 

(a) HCl0 2 {aq) 

(b) H 2 C0 3 (aq) 

(c) formic acid 

(d) perchloric acid 

5. Which one of the following compounds is not a strong electrolyte? 

(a) CH 3 COOH(aq) 

(b) HCl0 4 {aq) 

(c) HI(aq) 

(d) NaOH(aq) 

6. Which of the following is usually referred to as strong acid in water solution? Write the ionization 
reactions. 

(a) HF 

(b) HN0 2 

(c) H 2 C0 3 

(d) HS0 4 " 

(e) //7V0 3 

(f) //C/O4 

Further Reading / Supplemental Links 

• http://en.wikipedia.org/wiki 

Vocabulary 

strong acid Acids that completely ionize or undergo 100% ionization in solution (i.e. HCL). 

weak acids Acids that do not completely ionize or undergo 100% ionization in solution (i.e. HC 2 H 3 2 ). 

17.3 Arrhenius Bases 

Lesson Objectives 

• Define an Arrhenius base and know some examples of bases. 

• State the properties of bases. 

• Describe the neutralization reaction that bases undergo. 

Introduction 

Arrhenius broke ground in our understanding of acids and bases. He was the first to provide us with a 
definition from which we could identify an acid from a base. We have learned earlier the definition of an 

489 www.ckl2.org 



acid and now we will extend this definition to include bases. Keep in mind that Arrhenius came up with 
these theories in the late 1800s so his definitions came with some limitations. These limitations will be 
expanded upon later on in the unit. For now we will focus on his definitions. 

Bases Release OH-in Solution 

Arrhenius defined an acid as a substance that releases H + ions in solution. In contrast, he defined a base 
as a substance that releases OH ions in solution. Bases are ionic substances made up of a cation and an 
anion, of which the anion is the OH~ ion. It should be noted that very few of the hydroxides are actually 
soluble. If you recall from a previous chapter, only the alkali metals and Ba 2+ ions are soluble. Therefore, 
few of the basic solids will result in solutions when dissolved in water. One of the metal hydroxides that 
is very soluble is NaOH. The dissolving equation for NaOH is shown in Equation 1. 

NaOH^ -* Na + ( aq ) + OH~u^\ (Equation 1) 

Barium hydroxide produces a similar reaction when dissociating in water (Equation 2). 
Ba(OH) 2{s) -» Ba 2+ {aq) + 2 OH~ {aq) (Equation 2) 

The production of OH" ions is part of the conceptual definition of bases according to the Arrhenius 
definition of bases. 

Sample question: Write the chemical equation for the reaction of the following bases in water. 

(a) Lithium hydroxide (LiOH) 

(b) Sodium hydroxide (NaOH) 

(c) Potassium hydroxide (KOH) 
Solutions: 

(a) LiOH (s) -> Li + {aq) + OH {aq) 

(b) NaOH {s) -> Na + {aq) + OH~ {aq) 

(c) KOH (l) -> K+ {aq) + OH- {aq) 

Properties of Bases 

There is one common base that some may have had the opportunity to taste: milk of magnesia is a slightly 
soluble solution of magnesium hydroxide. This substance is used for acid indigestion. Flavorings have been 
added to improve the taste, otherwise it would have a bitter taste when you drink it. Other common bases 
include substances like Windex, Drano, oven cleaner, soaps and many cleaning other products. Please 
note: do not taste any of these substances. A bitter taste is one property you will have to take for granted. 

If you notice, one of the common bases is soap. This is an interesting example because another property 
of bases is that they are slippery to the touch. When you are washing dishes, you have probably noticed 
that the soapy water gets quite slippery. In fact, there may have been times when you dropped a dish or a 
glass or a cup because it was too slippery to hold on to. Now you know it isn't your fault - it's the soap's 
fault! 

As with acids, bases have properties that allow us to identify them from other substances. We have learned 
that acids turn blue litmus paper red. It stands to reason then that bases would turn red litmus paper 
blue (Figure ??). Notice that the effect of the indicator is the opposite of that of acids. 

www.ckl2.org 490 




As mentioned previously, bases have the ability to neutralize acids. Take for example the reaction between 
magnesium hydroxide (milk of magnesia) and HC1 (stomach acid) shown in Equation 3. 



Mg(OH) 2{s) + 2 HCl {aq) -» MgCl 2(aq) + 2 H 2 (L) 



(Equation 3) 



You see in Equation 3 that here again we have an acid and a base producing a salt and water. This is 
a similar reaction to the reaction between sodium hydroxide and hydrochloric acid. We will learn more 
about these reactions in the next section as well. 

Sample question: Write the balanced neutralization reaction between the following acids and bases. 

(a) HC10 4 + LiOH 

(b) HNO3 + Ba{OH) 2 

(c) H 2 S0 4 + KOH 
Solution: 

(a) HClO^aq) + LiOH {aq) -> LiCl0^ aq) + H 2 0( L ) 

(b) 2 HN0 3{aq) + Ba{OH) 2{aq) -> Ba(N0 3 ) 2{aq) + 2 H 2 {L) 

(c) H 2 S0 4 ( aq ) + 2 KOH( a ^ -> KiSO^aq} + 2 H 2 0^ 



Common Bases 



We mentioned above that bases are common to our everyday lives. If you do any of the cleaning in 
your home, you would use bases quite frequently. Drano©, used to unclog drains, is a solution of sodium 
hydroxide; sodium hydroxide, NaOH, is also used to make some soap; soft soap is often prepared with 
potassium hydroxide, KOH. We have already mentioned that magnesium hydroxide, Mg{OH) 2 , is used to 
make milk of magnesia. Windex is a water solution containing ammonium hydroxide, NH4OH. 

Soaps, as we know, have the ability to dissolve in water but also dissolve oil substances as well. The reason 
soap is able to do this is because of its structure. Soaps have long chains of carbon atoms, which make 
that part of the molecule nonpolar. This allows it to dissolve other nonpolar substances such as oils. And 
then on one end of the molecule is a sodium ion which makes that part of the molecule ionic. This ionic 
end of the molecule will be attracted to polar water molecules. This nonpolar-ionic molecule can attach 
any nonpolar particles to water molecules, and wash them away. 



491 



www.cki2.0rg 



CH 3 (CH 2 ) 16 



Hydrocarbon end Ionic end 

(Water repelling) (Water attracting) 




The Structure of Soap. 



Lesson Summary 

• An Arrhenius base is a substance that releases OH~ ions in solution. Bases turn red litmus paper 
blue, have a bitter taste, are slippery to the touch, and can neutralize acids. 

• A neutralization reaction between an acid and a base will produce a salt and water. Example: 
Mg(OH) 2{aq) + 2 HCl {aq) -* MgCl 2{aq) + 2 H 2 (L) . 

• Common bases include soap (NaOH) and Drano (NaOH), soft soap (KOH), milk of magnesia 
(Mg(OH) 2 ) and Windex (NH 4 OH). 

Review Questions 

1 . What is the role of litmus paper for acids and base chemistry? 

2. What are the properties of bases? Give a common example. 

3. Which statement best describes a characteristic of a base solutions? 

(a) They taste bitter. 

(b) They turn red litmus paper red. 

(c) They react with some metals to form hydrogen gas. 

(d) They are weak electrolytes. 

4. Which of the following is the Arrhenius definition of a base? 

(a) A base is a substance that donates protons. 

(b) A base is a substance that accepts protons. 

(c) A base is a substance that dissolves in water to form OH~ ions. 

(d) A base is a substance that reacts with water to form H + ions. 

5. Which of the following bases would be a weak electrolyte? 

(a) NaOH 

(b) Ba(OH) 2 

(c) Ca(OH) 2 

(d) Al(OH) 3 

6. Write the balanced neutralization reaction between the following acids and bases. 

(a) potassium hydroxide + hypochlorous acid 

(b) hydrobromic acid + calcium hydroxide 

(c) hydrochloric acid + sodium hydroxide 

(d) potassium hydroxide + sulfuric acid 

7. Write the net ionic equation for each of the neutralizations reactions in #4. 

Further Reading / Supplemental Links 

• http://en.wikipedia.org/wiki 
www.ckl2.org 492 



Vocabulary 

Arrhenius base A substance that produces OH" ions in a solution. 

17.4 Salts 

Lesson Objectives 

• Describe the formation of salts in neutralization reactions in terms of Arrhenius theory. 

• Identify acidic, basic, and neutral salts from neutralization reaction. 

Introduction 

Neutralization is a reaction between an acid and a base which produces water and a salt. The general 
reaction for the neutralization reaction is shown below. 

acid + base — » salt + water 

In this lesson, after reviewing the concept of ionic compounds, we will examine neutralization reactions 
and then look at the different type of salts that can be formed from acids and bases as they react in 
neutralization reactions. 

Ionic Compounds 

Ionic compounds are those formed between metal cations and nonmetal anions. When you take a look at 
the periodic table, the compounds with the most ionic character are those that are at opposite sides of the 
table. For example, sodium chloride will have more ionic character than zinc chloride. Look at Figure 
17.2 to see the positions of sodium, Na, zinc, Zn, and chlorine, CI. 



1 


1 


2 




::•■ 


i- 


1: 


M 


r 


is 


2 




















3 


Ma 












CI 




4 
























;.. 














* 






































: 






































- 







































Figure 17.2: The positions of sodium, zinc, and chlorine in the periodic table. 

When ionic compounds form between metal cations and nonmetal anions they transfer electrons and form 
charged particles and are capable of forming electrolyte solutions. If you notice, this is all similar to the 
descriptions of acids and bases. Acids are a combination of a hydrogen cation and a nonmetal anion. 
Examples include HCl, HNO3, and //C2//3O2. Bases can be a combination of metal cations and nonmetal 
anions. Examples include NaOH, KOH, and Mg{OH)2- 



493 



www.cki2.0rg 



Salts 

Salts that are ionic compounds result when an acid and a base react in a neutralization reaction. 

Acid + Base -» Salt + Water 

It is this salt that is formed that is the ionic compound. Let us examine some examples of the reactions 
we have looked at to see some of the salts and their ionic character. Take a look at Equations 1,2, and 3. 
Each of these equations has an acid and a base producing a salt and water. 

HCl {aq) + LiOH {aq) -> LiC\ aq) + H 2 {L) (Equation 1) 

2 HBr {aq) + Ba(OH) 2{aq) -> BaBr 2{aq) + 2 H 2 (L) (Equation 2) 

2 HF (aq) + Mg(OH) 2(aq) -» MgF 2{aq) + 2 H 2 {L) (Equation 3) 

Notice that in Equation 1, lithium chloride (LiCl) shows Li + ions and the Cl~ ions are the furthest away 
from each other on the periodic table so, generalizing, they will produce the salt with the most ionic 
character. Magnesium fluoride in Equation 3 has Mg 2+ ions and F— ion, which are closer to each other on 
the Periodic Table and therefore will have a little bit less ionic character. 

All acid-base reactions produce salts. According to the conceptual definition of the Arrhenius acid and 
base, the acid will contribute the H + ion that will react to neutralize the OH~ ion, contributed by the 
base, to produce neutral water molecules. The anion from the acid will combine with the cation from the 
base to form the ionic salt. Look at Equations 4 and 5 below to see the conceptual definition in action. 

HClO^aq) + NaOH( aq) -> NaCl0 4 ^ aq) + HOH^ L) (Equation 4) 

H 2 S <aq) + 2KOH {aq) -> K 2 S 4{aq) + 2 HOH (L) (Equation 5) 

(Note: HOH = H 2 0) 

No matter what the acid or the base may be, the products of this type of reaction will always be a salt 
and water. Aside from the fact that Arrhenius said that the H + ion will neutralize the OH~ ion to form 
water, we also know that these reactions are double displacement reactions and will therefore have their 
cations exchanging anions. 

To write this with a total ionic equation: 

H+ {aq) + Cl04~(aq) + Na+ \aq) + OH ( my ) -» Na + \ aq} + CIO4 \ aq) + H 2 0( L ) 
Or, since Na + and ClO^~ are spectator ions, the net ionic equation is: 

H+ {aq) + OH ' ( aq) -> H 2 0(l) 

This is the net ionic equation for all neutralization reactions. 
www.ckl2.org 494 



The Hydrolysis of Salts 
The Hydrolysis Reaction 

When a salt is dissolved in water, it is possible for the solution to be neutral, acidic, or basic. If a solution 
is to be acidic, it must contain more hydrogen ions than hydroxide ions and for the solution to be basic, it 
must contain more hydroxide ions than hydrogen ions. The reason that a dissolving salt may produce non- 
neutral solutions is that the cation or anion from the salt may not dissociate 100% when in the presence 
of hydrogen or hydroxide ions. Water, itself, dissociates very slightly into hydrogen and hydroxide ions. 
Therefore, when a salt is dissolved in water, there are some hydrogen and hydroxide ions available in the 
solution. Consider, first, the solution produced when the salt KBr dissolves in water. There will be four 
ions present in the solution. 

K + + Br + H + + OH ±5? 

When potassium ions in the solution come into contact with hydroxide ions, if the ions were to join together, 
the molecule formed would be KOH, which is a strong base, and would, therefore, immediately dissociate 
back into ions. Similarly, if the bromide ions come into contact with hydrogen ions, the molecule formed 
would be HBr, which is a strong acid, and would immediately dissociate back into ions. Having sodium 
and chloride ions in a water solution would not cause a reaction. 

Consider, now, the solution produced when the salt NH4CI is dissolved in water. There will be four ions 
present in the solution. 

NH+ + CI + H + + OH ±5? 

When hydrogen ions come into contact with chloride ions, if they join together, the resultant molecule 
would be HCl, which is a strong acid, and therefore the HCl would immediately dissociate back into 
the ions. When NH^ + ions come into contact with OH ions, however, the resultant molecule would be 
NH4OH, which is a weak base and therefore, does not dissociate very much. Therefore, when ammonium 
chloride is dissolved in water, a reaction occurs. 

NH 4 + + Cr + H + + OH ±* NH A OH {aq) + Cl~ + H + 

The ammonium hydroxide dissociates very little, so we would have mostly un-dissociated ammonium 
hydroxide molecules in solution with hydrogen and chloride ions. The hydrogen ions in this final solution 
would cause the solution to be acidic. Thus, dissolving ammonium chloride in water produces an acidic 
solution. 

By a similar process, dissolving sodium acetate, NaC 2 H 3 2 , in water will produce a basic solution. When 
the sodium acetate is dissolved in water, four ions will be present in the solution. 

Na + + C 2 H 3 2 +H+ + OH £?? 

If sodium ions contact hydroxide ions, the substance formed would be a strong base which would immedi- 
ately dissociate. If hydrogen ions contact acetate ions, however, the molecule formed would be acetic acid, 
which is a weak acid and the ions would NOT dissociate. Therefore, when sodium acetate is dissolved in 
water, a reaction will occur as shown below. 

Na + + C 2 H 3 2 + H + + OH £5 HC 2 H 3 2{aq) + Na + + OH 

The resultant solution has excess hydroxide ions and therefore, is basic. The dissolving of sodium acetate 
salt in water produces a basic solution. 

The reactions between some of the ions in the salts and water are called hydrolysis reactions. 

495 www.ckl2.org 



Neutral, Acidic, and Basic Salts 

We have talked a little already about the fact that there are not just strong acids but there are also weak 
acids as well. The difference between them is the percent ionization. The same is true for bases. Strong 
bases dissociate 100% and weak bases do not. Table 17.2 shows all of the strong acids and bases, all of 
the rest of the acids and bases are weak. 

Table 17.2: Strong Acids and Bases 



Strong Acid 



Formula 



Strong Base 


Formula 


Lithium hydroxide 


LiOH 


Sodium hydroxide 


NaOH 


Potassium hydroxide 


KOH 


Rubidium hydroxide 


RbOH 


Cesium hydroxide 


CsOH 


Calcium hydroxide 


Ca(OH) 2 


Strontium hydroxide 


Sr{OH) 2 


Barium hydroxide 


Ba(OH) 2 



Hydrochloric Acid 


HCl 


Hydrobromic Acid 


HBr 


Hydroiodic Acid 


HI 


Nitric Acid 


HN0 3 


Perchloric Acid 


HClOi 


Sulfuric Acid 


H 2 S0 4 



So why is this important to us here? The information in the table helps us to determine what type of salt 
is formed in an acid-base reaction. For example if we have a reaction between a strong acid and a strong 
base we form a neutral salt. It is like a power struggle between the acid and the base. Since both are 
strong, there is no compound with more power and thus the salt ends up being neutral. If, however, we 
have a reaction between a weak acid and a strong base (Equation 6), the result would be a basic salt. 



HC 2 H 3 2 (aq) 


+ 


NaOH (aq) 


-> 


NaC 2 H 3 2 (aq) 


+ 


#20(X) 


(Equation 6) 


Acetic acid 




sodium hydroxide 




sodium acetate 




water 




(weak acid) 




(strong base) 




(basic salt) 









When the basic salt is dissolved in water, a reaction called hydrolysis takes places in which extra hydroxide 
ions, OH 1 —, are produced from the salt and the water molecules. Since the hydroxide ions come from a 
salt, it is called a basic salt. 

A similar situation will occur when we have a strong acid reacting with a weak base. When a strong acid 
reacts with a weak base, an acidic salt is formed. 



HCl( aq ) 

Hydrochloric acid 
(strong acid) 



+ NH 4 OH {aq) 

ammoniun hydroxide 
(weak base) 



NHiCl^q) + H 2 {L) (Equation 7) 

ammonium chloride water 

(acidic salt) 



Again, a hydrolysis reaction takes place. The salt will react with water molecules and produce excess 
hydrogen ions, H + , in solution and is therefore referred to as an acidic salt. 

Sample question: Complete the following neutralization reactions and identify the type of salt produced. 

(a) H 2 S0 4(ag) + Ba(OH) 2{aq) -»? 

(b) HCOOH {aq) + Ca(OH) 2 (aq) -»? 
Solution: 

(a) H 2 S0 4{aq) + Ba{OH) 2{aq) -» BaSO i{aq) + 2H 2 {L) 



www.ckl2.org 



496 



Strong acid + strong base = neutral salt 

(b) 2 HCOOH (aq) + Ca{OH) 2(aq) -» Ca(HCOO) 2{aq) + 2 Zf 2 (L) 

Weak acid + strong base = basic salt 

It can also be determined which acid and base was used to form the salt and from this you could determine 
if your starting reaction was a strong acid or a weak acid, a strong base or a weak base. Watch how this 
works. You were given that the product salt was calcium nitrate, Ca(NOs)2- Remember that there is a 
double displacement reaction that forms the salt so we can write parts of the reaction: 



H? + ?0H *■ Ca(N0 3 ) 2 + HOH 




Calcium must come from the Nitrate must come from the 

base. Therefore, the base acid. Therefore, the acid 

must be Ca(0H} 2 . mustbeHNQ s . 

Ca(0H) 2 is a strong base. HN0 3 is a strong acid. 

Therefore, the neutralization reaction would have been: 

2 HN0 3 + Ca(OH) 2 -» Ca(N0 3 ) 2 + 2 HOH (Equation 8) 

So this salt would have been produced from a strong acid-strong base reaction! 

Let's try another. Consider the salt copper (II) chloride (CuCl 2 ). 

The copper would have come from the base, Cu(OH) 2 , which is a weak base. The chloride would have 
come from the acid, HCl, which is a strong acid. 

2 HCl + Cu(OH) 2 -> CuCl 2 + 2 HOH (Equation 9) 

Therefore the reaction is a strong acid weak base reaction! 

Lesson Summary 

• Ionic compounds are those formed between metal cations and nonmetal anions. Acids are a com- 
bination of hydrogen and a nonmetal anion. Bases can be a combination of metal cations and 
nonmetal anions. Ionic compounds result when an acid and a base react in a neutralization reaction 
(Acid+ Base -> Salt + Water). 

• According to the conceptual definition of the Arrhenius acid and base, the acid will produce the H + 
ion, which will react to neutralize the OH' ion produced by the base to produce the neutral water. 
The other product will be the ionic salt. A strong acid + a strong base in an acid/base neutralization 
reaction will form a neutral salt. 

• A strong acid + a weak base in an acid/base neutralization reaction will form an acidic salt. A weak 
acid + a strong base in an acid/base neutralization reaction will form a basic salt. 

Review Questions 

1. How do an acid and a base fit the definition of an ionic compound? Use examples in your answer. 

497 www.ckI2.org 



2. Explain neutralization reactions in terms of Arrhenius theory. Use an example in your answer. 

3. Which salt will form a basic solution when dissolved in water? 

(a) KN0 3 

(b) CaCl 2 

(c) NaClOi 

(d) NaN0 2 

4. Which salt will form an acidic solution when dissolved in water? 

(a) copper(II) sulfate 

(b) sodium acetate 

(c) potassium chloride 

(d) sodium cyanide 

5. Milk of magnesia is a common over- the - counter antacid that has, as its main ingredient, magnesium 
hydroxide. It is used by the public to relieve acid indigestion. Acid indigestion is caused by excess 
stomach acid being present. Since a stomach upset is caused by excess hydrochloric acid, this tends 
to be a quite painful affliction for people. Write the balanced chemical equation for the reaction 
between milk of magnesia and hydrochloric acid. What type of reaction is this? What type of salt is 
formed? 

6. Complete the following neutralization reactions and identify the type of salt produced. 

(a) H2S Oi(aq) and NaOH{aq) — > 

(b) HN0 3 (aq) and NH A OH{aq) -> 

(c) HF(aq) and NH A OH{aq) -> 

(d) CH 3 COOH(aq) and KOH{aq) -> 

(e) HCl(aq) and KOH(aq) -> 

Further Reading / Supplemental Links 

• http : //en . wikipedia . org/wiki/Salts 

Vocabulary 

basic salt A salt formed in a neutralization reaction between a weak acid and a strong base. 

acidic salt A salt formed in a neutralization reaction between a strong acid and a weak base. 

neutral salt A salt formed in a neutralization reaction between a strong acid and a strong base or a 
weak acid and a weak base. 

Labs and Demonstrations for Acids and Bases 
Teacher's Pages for Hydrolysis of Salts 



A 



Investigation and Experimentation Objectives 



In this activity, the student will predict the acidity or basicity of hydrated salts and then carry out 
laboratory activities to compare the relationships between evidence and explanations. 

www.ckl2.org 498 



Lab Notes 

Preparation of Solutions: 

200 mL of each solution should be more than enough to complete this lab. To prepare each solution, mass 
the specified amount of reagent, dissolve it in 150 mL of water, and dilute the resulting solution to 200 mL: 

CuSOi- 5H2O : 5.0 grams 

Ca(NOs)2 : 3.3 grams 

K3PO4 : 4.2 grams 

KCl : 1.5 grams 

NaBr : 2.1 grams 

Na2S : 1.6 grams 

(NH 4 ) 2 C0 3 : 1.9 grams 

Na<iCrOt± : 3.2 grams 

MgBr 2 : 3.7 grams 

NaCl : 1.2 grams 

Use care when opening the container of (NH 4 )2C0 3 . It undergoes decomposition over time, and outgases 
NH 3 , which collects in the container. Upon opening the NH 3 diffuses out rapidly, and is very irritating to 
eyes and skin. Live and learn. 

Answers to Pre-Lab Questions 

1. Write dissociation equations for the following salts: 



(a 
(b 

(c 
(d 

(0 
(f 
(g 
(li 
(i 
(j 



CuS0 4 ^Cu 2+ + SO\ + 
Ca{N0 3 ) 2 -» Ca 2+ + 2 N0 3 
K 3 P0 4 -> 3 K+ + POf- 
KCl -> K + + Ct 
KBr -» K + + Br 
Na 2 S -^2 Na + + S 2 - 
(NH 4 ) 2 C0 3 -> 2 NH+ + COf- 
Na 2 CrOi -^ 2 Na + + CrOf 
MgBr 2 -» Mg 2+ + 2 Br 
NaCl -^ Na + + CI 



2. NaC e H 5 C0 2 + H 2 -^ HC 6 H 5 C0 2 + Na + + OH solution will be basic 

Lab — Hydrolysis of Salts 

Background Information 

A salt is an ionic compound containing positive ions other than hydrogen and negative ions other than 
hydroxide. Most salts will dissociate to some degree when placed in water. In many cases, ions from the 
salt will react with water molecules to produce hydrogen ions, H + , or hydroxide ions, OH . Any chemical 
reaction in which water is one of the reactants is called a hydrolysis reaction. Salts are usually formed from 
the neutralization reaction between an acid and a base. A salt formed from a strong acid and a strong base 
will not undergo hydrolysis. The resulting solution is neutral. An example of such a salt is KBr, formed 
from a strong acid, HBr, and a strong base, KOH. 

Salts formed from the reaction of a strong acid and a weak base hydrolyzes to form a solution that is 

499 www.ckl2.org 



slightly acidic. In this kind of hydrolysis, the water molecules actually react with the cation from the weak 
base. For example, when ammonium chloride, NH4CI, hydrolyzes, water molecules react with the NHf 
ion: 

NH+ + H 2 -> NH4OH + H + 

The formation of the H + ion from this reaction makes the solution acidic. 

Salts formed from the reaction of a weak acid and a strong base hydrolyze to form a solution that is slightly 
basic. In this kind of hydrolysis, it is the anion from the weak acid that actually reacts with the water. 
For example, when sodium acetate, NaC 2 H 3 2 , hydrolyzes, water molecules react with the acetate ion: 

C 2 H 3 2 +H 2 -> HC 2 H 3 2 + OH 

The formation of the OH ion from this reaction makes the solution basic. Salts formed from a weak acid 
and weak base produce solutions that may be slightly acidic, slightly basic, or neutral, depending on how 
strongly the ions of the salt are hydrolyzed. 

In this experiment you will test several different salt solutions with pH paper and phenolphthalein solution 
to determine their acidity or basicity. 

Purpose 

To determine the relative acidity or basicity of various salt solutions, and thus predict whether hydrolysis 
occurred, and if so, what the reaction products are. 

Pre-Lab Questions 

1. Write dissociation equations for the following salts: 



(a 
(b 

(c 
(d 

(e 

(f 



Copper(II) sulfate 
Calcium nitrate 
Potassium phosphate 
Potassium chloride 
Potassium bromide 
Sodium sulfide 
Ammonium carbonate 
Sodium chromate 
Magnesium bromide 
Sodium chloride 



(h) 
(i) 
0) 

2. Sodium benzoate is the salt formed in the neutralization of benzoic acid with sodium hydroxide. 
Benzoic acid is a weak acid. Write the hydrolysis reaction for the dissolution of solid sodium benzoate, 
NaC^H^C0 2 , in water. Will sodium benzoate solution be acidic, basic, or neutral? 

Apparatus and Materials 

• 10 small or medium sized test tubes, or a micro reaction plate 

• Test tube rack 

• 0.1 M solutions of cupric sulfate, calcium nitrate, potassium phosphate, potassium chloride, sodium 
bromide, sodium sulfide, ammonium carbonate, sodium chromate, magnesium bromide and sodium 
chloride 

• Universal pH indicator paper, range 0-14 

• Phenolphthalein indicator solution 

www.ckl2.org 500 



• 10 mL graduate 

• Stirring rod 

Safety Issues 

The solutions used may be slightly acidic or basic, and as a result can be corrosive or caustic. Use proper 
laboratory safety equipment and techniques. 

Procedure 

1. Obtain a clean, dry micro reaction plate, or 10 test tubes 

2. To test tubes 1 through 10 or the reaction plate, add eight to ten drops of the following solutions: 

Tube or Well 1: Cupric sulfate 
Tube or Well 2: Calcium nitrate 
Tube or Well 3: Potassium phosphate 
Tube or Well 4: Potassium chloride 
Tube or Well 5: Sodium bromide 
Tube or Well 6: Sodium sulfide 
Tube or Well 7: Ammonium carbonate 
Tube or Well 8: Sodium chromate 
Tube or Well 9: Magnesium bromide 
Tube or Well 10: Sodium chloride 

3. Add two drops of phenolphthalein solution to each of the occupied wells of the microplate or test tube. 
Record your observations in the data table. 

4. Test each solution with pH paper and record your results. 
Data 

Table 17.3: Data Table 
Well # Salt 



1 
2 
3 
4 
5 
6 
7 
8 
9 
10 



Post-Lab Questions 

1. Where a hydrolysis is likely to occur in each of the following, write a net ionic hydrolysis equation. If 
no hydrolysis is likely, write NR. 



Effect on 


pH 


Original 


Strong 


Original 


Strong 


Indicator 




Acid 


or Weak 
Acid 


Base 


or Weak 
Base 



501 www.ckl2.org 



Cu 2+ + SOf + 2 Zf 2 0-» 

Ca 2+ + 2 N<9 3 + 2 H 2 -> 

3 K + + POf + 3 H 2 -> 

k + + cr + H 2 -> 

Ma + + Br + # 2 <9 -> 

2 Na + + S 2 ~ + H 2 -> 

2 JV/f+ + CO 2 ^ + Zf 2 -» 

2 Na + + CrOf" + 2 Zf 2 -» 

Mg 2+ + 2 5r" + 2 H 2 -» 

Na + + C7 - + Z/ 2 -» 

2. How do your observations and /?// readings compare with the expected results based on the equations 
for the hydrolysis reactions? 

3. What is a spectator ion? Name the spectator ions present in each hydrolysis reaction in this experiment. 

4. A salt formed from a strong acid and a strong base produces a neutral solution. A salt of a weak acid 
and a weak base may or may not produce a neutral solution. Explain why. 

5. Bases make effective cleaning agents, because they can convert grease and oils to a water soluble 
substance. Trisodium phosphate (TSP) is a common commercially available cleaner. Give the reaction 
TSP undergoes to create a basic solution. 

17.5 pH 

Lesson Objectives 

• Calculate [H + ] for strong acids and [OH~] for strong bases. 

• Define autoionization and use it to find [H + ] from [OH~] or to find [OH~] from [H + ]. 

• Describe the pH scale. 

• Define pH. 

• Calculate pH from [H + ] or vice versa. 

Introduction 

A few very concentrated acid and base solutions are used in industrial chemistry and occasionally in 
inorganic laboratory situations. For the most part, however, acid and base solutions that occur in nature, 
those used in cleaning, and those used in organic or biochemistry applications are relatively dilute. Most 
of the acids and bases dealt with in laboratory situations have hydrogen ion concentrations between 1.0 M 
and 1.0 x 10~ 14 M. 

Expressing hydrogen ion concentrations in exponential numbers becomes tedious and is difficult for those 
not trained in chemistry. A Danish chemist named Soren Sorenson developed a shorter method for express- 
ing acid strength or hydrogen ion concentration with a non-exponential number. He named his method 
pH and while the exact definition of pH has changed over the years, the name has remained. 

pH, today, is defined as the negative logarithm of the hydrogen ion concentration. 
www.ckl2.org 502 



pH = -log [H + ] 

If the hydrogen ion concentration is between 1.0 M and 1.0 x 10 -14 M, the value of the pH will be between 
and 14. 



Concentrations of Ions 

We have said that a strong acid completely ionizes in solution and therefore in solution there are no intact 
acid molecules remaining but only the H + cations and the anions from the acid. Look at Equation 1 as an 
example. 

HC\ aq) -> H + {aq) + Cr {aq) (Equation 1) 

100% ionized 

We can extend this concept to give actual concentrations of the hydronium ion for strong acid solutions. 
If we had a 0.10 mol/L solution of hydrochloric acid and it completely ionizes, then the concentration of 
H + ions that are produced are also 0.10 mol/L. Look at the balancing coefficients for Equation 1. The 
balancing coefficients are all ones (1). Therefore the concentrations will all be the following: 

HCl {aq) -> H+ {aq) + Cl ~(aq) (Equation 2) 

0.10 mol/L 0.10 mol/L 0.10 mol/L 

For strong bases, the same calculation can be performed. Since strong bases are 100% dissociated, when 
we are given the concentration of the strong base we can then conclude the concentration of the hydroxide 
ion. Let's look at the following example. What would be the [OH~] for a solution of Ba(OH)2, knowing 
the concentration of Ba{OH)2 is 0.24 mol/L? 

Ba(OH) 2{aq) -> Ba 2+ {aq) + 2 OH~ {aq) (Equation 3) 

100% dissociated 

Notice both the 100% dissociation and the balancing coefficients. Now let's fill in the coefficients. Take 
a look at Equation 4 to see how the balancing coefficients have helped determine the [OH~] in the same 
manner as they did for finding the [H + ]. 

Ba{OH) 2{aq) -> Ba 2+ {aq) + 20H~ {aq) (Equation 4) 

0.24 mol/L 0.24 mol/L 0.48 mol/L 

Sample question: What would the [H + ] be for the ionization of a 0.35 mol/L solution of nitric acid? 
Solution: 

HN0 3{aq) -> H + {aq) + N0 3 - {aq) 

0.35 mol/L 0.35 mol/L 0.35 mol/L 

503 www.ckl2.org 



Relationship Between [H+] and [OH-] 



Concentrated hydrochloric acid is found in most high school chemistry labs. This acid is 12.0 mol/L or 
12 mol of HCl dissolved in enough water to make 1 L. It is the presence of the water molecules that 
allows that hydrogen chloride to ionize and produce the hydrogen ions that end up in solution. Water 
also undergoes ionization. Actually, water undergoes autoionization. Autoionization is when the same 
reactant acts as both the acid and the base. Look at the reaction below. 

#2#(L) + #2<9(L) ^> H 3 + ( aq } + OH'fa) 

From experimentation, chemists have determined that in pure water [H + ] = 1 x 10~ 7 mol/L and [0// _1 ] = 
1 X 1CT 7 mol/L. In other words, //20(n is autoionizing and so the [H + ] = [OH"] = 1 x 1CT 7 mol/L. 

The ionization of water is frequently written as H20n\ t> H + + OH 

The equilibrium constant expression for this dissociation would be: 

K w = [H+][OH-} 

Like other equilibrium constants that are for special reactions, this K is given its own subscript. Since this 
equilibrium constant is for the dissociation of water, it is designated as K w . The ion product constant for 
water, K w is the product of the hydronium ion and the hydroxide ion concentrations in the autoionization 
of water. 

We can the calculate K w because we know the value of [H + ] and [OH"] for pure water at 25° C. 

K w = [H+][OH-] 

K w = (lxl(T 7 )(lxl(r 7 ) 

K w = 1 x 1(T 14 

A further definition of acids and bases can now be made: 
when 

[//3<9 + ] = [OH~l] (as in water), the solution is neutral 
[H%0 + ] > [OH" I] the solution is an acid 
[//3<9 + ] < [0// _ l] the solution is a base 

To continue with these ideas: 

an acid has a [H%0 + ] that is greater than 1 x 1CT 7 and the [OH" 1 ] is less than 1 x 1CT 7 

a base has a [O// 1- ] that is greater than 1 x 1CT 7 and the [//30 + ] is less than 1 x 10~ 7 

The equilibrium between the water molecules and the H + and OH" ions will exist in all water solutions 
regardless of anything else that may be present in the solution. Some of the other substances that are 
placed in water solution may become involved with either the hydrogen or hydroxide ions and alter the 
equilibrium state. In all cases, as long as the temperature is 25°C, the equilibrium, H<iOn\ ^ H + + OH~ , 
will shift and maintain the equilibrium constant, K w , at exactly 1 x 1CT 14 . 

For example, a sample of pure water at 25°C, has a [H + ] equal to 1 x 10~ 7 M and a [OH"] = 1 x 10^ 7 M. 
The K w for this solution, of course, will be 1 x 10~ 14 . Suppose some HCl gas is added to this solution so 
that the H + concentration increases. This is a stress to the equilibrium condition. Since the concentration 
of a product is increased, the reverse reaction rate will increase and the equilibrium will shift toward the 

www.ckl2.org 504 



reactants. The concentrations of both ions will be reduced until equilibrium is re-established. If the final 
[H + ] — 1 X 10~ 4 M, we can calculate the [0H~] because we know that the product of [H + ] and [0H~] at 
equilibrium is always 1 x 10 -14 . 



K w = [H + ][OH-] = lxl(T 14 

. lxl(T 14 lxlO~ 14 in 

OH-} = r - = r = 1 x 10" 10 M 

L J [H+] lxlO" 4 



Suppose, on the other hand, something is added to the solution that reduces the hydrogen ion concentration. 
As soon as the hydrogen ion concentration begins to decrease, the reverse rate decreases and the forward 
rate will shift the equilibrium toward the products. The concentrations of both ions will be increased until 
equilibrium is re-established. If the final hydrogen ion concentration is 1 x 10~ 12 M, we can calculate the 
final hydroxide ion concentration. 



Kw — 



[H + ][OH-] = lxl(T 14 



, lxlO~ 14 lxlO -14 

\OH~\ = r n = tt = 1 x 10~2 M 

L J [H+] lxlO- 12 



Using the K2 expression and our knowledge of the K2 value, anytime we know either the [H + ] or the [OH ] 
in a water solution, we can always calculate the other one. 

Sample question: What would be the [H + ] for a grapefruit found to have a [OH~] of 1.26 x 1CT 11 mol/L? 
Is the solution acidic, basic, or neutral? 

Solution: 

K w = [H + ][OH~] = 1.00 x 10 -14 

r ,, 1.00 xlO -14 1.00 xlO -14 

\H + ] = — : : — = rr = 7.94 X 10" 4 M 

L J [OH-] 1.26 xlO- 11 

The pH Scale 
Introduction to pH 

The pH scale developed by S0rensen is a logarithmic scale. Not only is the pH scale a logarithmic scale 
but by defining the pH as the negative log of the hydrogen ion concentration, the numbers on the scale 
get smaller as the hydrogen ion concentration gets larger. For example, pH = 1 is a stronger acid than 
pH = 2 and, it is stronger by a factor of 10. A solution whose pH = 1 has a hydrogen ion concentration 
of 0.10 M while a solution whose pH = 2 has a hydrogen ion concentration of 0.010 M. You should note 
the relationship between 0.10 and 0.010, 0.10 is 10 times 0.010. This is a very important point when using 
the pH scale. 

As was pointed out earlier, when the [H^O + ] is greater than 1 x 10~ 7 , a solution is considered to be an acid. 
However, there are a great number of possibilities when you consider that an acid can have a hydrogen ion 
concentration that is anything greater than 1 x 10 -7 . The same can be seen to be true for bases - just the 
opposite direction (Figure 17.3). 

There are a couple of important points to note here. First, all the numbers on the pH scale represent 
numbers with negative exponents and exponential numbers with negative exponents are small numbers; 
so all of these numbers represent [H + ] that are less than one. Second, there are acids and bases whose H + 
concentrations do not fit within this range and therefore, not all acid or base strengths can be represented 
on the pH scale. 

505 www.ckl2.org 



The pH scale found in Figure 17.3 shows that acidic solutions have a pH within the range of up to but 
not including 7. The closer the pH is to the greater the concentration of H^O + ions and therefore the 
more acidic the solution. The basic solutions have a pH with the range from 7 to 14 (Table 17.4). The 
closer the pH is to 14, the higher the concentration of OH' ion and the stronger the base. For 25°C, a pH 
of 7 is neutral which means that [H 3 + ] = [OH~] = 1 x 1(T 7 M. 

Table 17.4: 

pH level Solution 

pH < 7 acidic 

pH = 7 neutral 

pH > 7 basic 



Acid Base 

Neutral 



I 



1 2 3 4 5 6 78 9 10 11 12 13 14 



Figure 17.3: The pH Scale. 



pH = -log [H+] 

S0rensen's idea that the pH would be a simpler number to deal with in terms of discussing acidity level 
led him to a formula that relates pH and \H + \. This formula is: 

pH = -log [H + ] 

where p = -log and H refers to the hydrogen ion concentration. The p from pH comes from the German 
word potenz meaning power or the exponent of. In this case the exponent is 10. Therefore, [H + ] = 10 _pW . 

When the [H+] = 0.01 mol/L, the pH will be 

pH = -log (0.01) 
pH = -log (1 x 10" 2 ) 
pH = 2 

Since we are talking about negative logarithms (-log), the more hydrogen ions that are in solution, the 
more acidic the solution and the lower the pH. 

The Mathematics of pH 

pH measures the level of acidity in solution within a certain range and has the definition, pH = -log [H + ]. 
pH is a logarithmic scale which means that a difference of 1 in pH units indicates a difference of a factor 
of 10 in the hydrogen ion concentrations. If we have the pH of a solution and are asked to find the [H + ], 
the formula for pH can be converted to a formula for [H + ] by taking the inverse log of both sides of the 
equation. That process yields: 

www.ckl2.org 506 



[H+] = 10 



pH 



Example 1: Determine the pH of a solution that has a [H + ] = 1 X 10 8 . 
Solution: 

pH = -log [H + ] = -log (1 x 10~ 8 ) 

The log of 1 is and the log of 10~ 8 is -8. 

pH = -(0 - 8) = 8 

Sample question 2: Calculate the [H + ] given that the pH is 4. 
Solution: 

[H+] = 10~ p// 

[//+] = 10~ 4 

[H + ] = 1 x 10~ 4 mol/L 

Sometimes you will need to use a calculator. 

Sample question 3: Calculate the pH of saliva with [H + ] = 1.58 x 10~ 6 mol/L. 

Solution: 

pH = -log [H + ] = -log (1.58 x 10~ 6 ) 
pH = 5.8 

Sample Question 4: Fill in the rest of the Table 17.5. 

pH = -log [H + ] 

Table 17.5: 

[H+] (mol/L) -log [H+] pH 

0.1 1.00 1.00 

0.2 0.70 0.70 

1.00 xl0~ 5 ? ? 

? ? 6.00 

0.065 ? ? 

? ? 9.00 



Solution: 



pH = -log [#+] 



50T www.ckl2.org 



Table 17.6: 



[H+] (mol/L) -log [H+] pH 

0.1 1.00 1.00 

0.2 0.70 0.70 

1.00 xl0~ 5 5.00 5.00 

1.00 xl0~ 6 6.00 6.00 

0.065 1.19 1.19 

1.00 xl0~ 9 9.00 9.00 

We said that the pH scale was one that showed the pH becoming lower as the strength of the acid becomes 
larger. Let's think about this for a second. Stomach acid is HCl, a strong acid. Strong acids are powerful, 
we can assume because they completely ionize and therefore would have all of their H + ions present in the 
solution when the reaction is complete. Vinegar (that we may put on our salad, cook with, make those 
neat science fair volcanoes with!) is a weak acid. It only partially ionizes and only allows some of its H + 
ions to come into solution. Therefore, the pH of HCl, according to this observation would be lower than 
that of vinegar. Look at Figure 2. What does the pH scale diagram tell us about the pH of 0.1 M HCl 
and 5% vinegar? Sure enough, the pH for HCl is 1.0 and that of 5% vinegar is around 2.8. 

Have you ever cut an onion and had your eyes water up? This is because of a compound with the formula 
C^HqOS that is found in onions. When you cut the onion, a variety of reactions occur that release a 
gas. This gas can diffuse into the air and eventfully mix with the water found in your eyes to produce a 
dilute solution of sulfuric acid. This is what irritates your eyes and causes them to water. There are many 
common examples of acids and bases in our everyday lives. Look at the pH scale (Figure 17.4) to see how 
these common examples relate in terms of their pH. 

Soda Pop 

Pure Water Soa P 

NaOH 



HCl Vinegar 



. I Tomatoes Milk Eqqs Detergent ' 

L J i i u ill 



i 



01 2 3.45 6 7T 89 10 12 13 14 



Blood . T 

I Ammonia 

Lemon Juice Milk of Magnesia 

Orange Juice 

Figure 17.4: pH Scale for Common Substances. 

You can see that pH definitely does play a role in our everyday lives as we come in contact with many 
substances that have varying degrees of acidity. We may not necessarily think about the pH of eggs as we 
eat them or think about how acidic orange juice is when we enjoy a glass, but maybe next time we have 
breakfast we may ponder for a second about the work of S0rensen ... just maybe! 

Summary 

• Water dissociates to a very slight degree according to the equation ^Ora ^=> [H + ] + [OH~]. 
. In pure water at 25°C, [H+] = [OH~] = 1.00 x 10~ 7 M. 

• The equilibrium constant for the dissociation of water, called K w , at 25°C is equal to 1.00 x 10 -14 . 

• The definition of pH is pH = -log [H + ]. 

www.ckl2.org 508 



Review Questions 

1. Why is it necessary to balance the chemical equation before determining the [H + ] or [0H~\ for strong 
acids and bases? 

2. Why can't you determine the [H + ] or [OH'] for weak acids and bases the same way you can determine 
the [H + ] or [0H~] for strong acids and bases? 

3. What is the [H + ] ion concentration in a solution of 0.350 mol/L H2SO4? 

(a) 0.175 mol/L 

(b) 0.350 mol/L 

(c) 0.700 mol/L 

(d) 1.42 x lO' 14 mol/L 

4. A solution has a pH of 6.54. What is the concentration of hydronium ions in the solution? 

(a) 2.88 x 10" 7 mol/L 

(b) 3.46 x 10" 8 mol/L 

(c) 6.54 mol/L 

(d) 7.46 mol/L 

5. A solution has a pH of 3.34. What is the concentration of hydroxide ions in the solution? 

(a) 4.57 x 10~ 4 mol/L 

(b) 2.19 x 10 -11 mol/L 

(c) 3.34 mol/L 

(d) 10.66 mol/L 

6. A solution contains 4.33 x 10~ 8 M hydroxide ions. What is the pH of the solution? 

(a) 4.33 

(b) 6.64 

(c) 7.36 

(d) 9.67 

7. Fill in the Table 17.7 and rank the solutions in terms of increasing acidity. 

Table 17.7: 

Solutions [H+] (mol/L) -log [H + ] pH 

A 0.25 0.60 0.60 

B ? 2.90 ? 

C 1.25 xl0~ 8 ? ? 

D 0.45 xl0~ 3 ? ? 

E ? 1.26 ? 



8. A bottle of calcium hydroxide is found in the lab with a label reading: 0.014 mol/L. 

(a) What are the concentrations of all of the ions present in the solution? 

(b) What is the pH of the solution? 

9. It has long been advocated that red wine is good for the heart. However, wine is also considered to 
be an acidic compound. Determine the concentration of hydronium ions in wine with pH3.81. 

10. The diagram that follows represents a weak acid before ionization and when the reaction comes to 
equilibrium. If the acid were weaker than the one represented in the diagram, how would the diagram 
change? Draw a new diagram to represent your answer. If the acid were a strong acid, how would 

509 www.ckl2.org 



the diagram change? Draw a new diagram to represent your answer. 



100 
90 f- 
80 -|~ 

70 



■ Weak Acid 

■ H* Ion 
□ Anion 



Equilibrium 



Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

pH scale A scale measuring the [H + ] with values from to 14. 

pH = -log [H + ] - Formula used to calculate the power of the hydronium ion. 

autoionization Autoionization is when the same reactant acts as both the acid and the base. 

ion product constant for water K w , is the product of the hydronium ion and the hydroxide ion con- 
centrations in the autoionization of water. 

17.6 Weak Acid/Base Equilibria 

Lesson Objectives 

• Define weak acids and weak bases in terms of equilibrium. 

• Define K a and Kb- 

• Use K a and K b to determine acid and base strength. 

• Use K a and Kb in acid/base equilibrium problems. 

Introduction 

Acid and base equilibria are another form of equilibrium reactions that deserve special mention. In prior 
learning we had discussed that weak acids and bases do not completely ionize or dissociate in solution. 
Thus, their reactions also possess an equilibrium state where their forward reaction rates equal their reverse 
reaction rates. Since this is the case, equilibrium constant expressions and equilibrium calculations can 
be completed for these reactions as was done for all other equilibria. When there are a group of reactions 
that are similar in appearance and function, the equilibrium constant for those reactions are often given a 
subscript denoting them as belonging to the group. Such was the case for solubility product constants which 
were given the designation K sp . The dissociation of weak acids are also a special group of reactions and 
their equilibrium constant is designated K a . The equilibrium constants for the dissociation of weak bases 
are designated Kb- In this section, we will focus our attention to a special case of equilibrium reactions, 
the acid - base equilibria. 



www.ckl2.org 



510 



Weak Acids and Weak Bases are Equilibrium Systems 

Strong acids are denned as acids that completely ionize in water solution; strong bases as those that 
completely dissociate in water solution. Look at Equation 1 and Equation 2 for an example of a strong 
acid ionization reaction and a strong base dissociation reaction, respectively. 

HClfa) -» H + {aq) + Cl~ {aq) (Equation 1) 

NaOH^ -> Na + ( a? ) + OH~ ^ (Equation 2) 

Weak acids and bases, however, do not completely ionize or dissociate in solution. In the chapter on weak 
acids and bases, we learned that solutions of weak acids and bases ionize or dissociate much less than 
100%. Ammonia is a weak base, for example. Look at Equation 3, the dissociation reaction for ammonia. 

NH 3{aq) + H 2 {L) £; NH 4 + {aq) + OH~ {aq) (Equation 3) 

Ammonia has an equilibrium constant, Kb, of 1.8 x 10~ 5 . This value is less than 1. This means that the 
equilibrium position favors the reactants and that there is not very much reaction of ammonia to form the 
protonated ammonium ion. Said another way, at equilibrium, there are many more ammonia molecules 
present in solution than there are ammonium ions or hydroxide ions. 

The same is true for weak acids. Vinegar is a common weak acid and represents an approximately 5% 
acetic acid solution. Equation 4 shows the ionization of acetic acid or vinegar. 

HC 2 H 3 2 (a q ) ^> H+ {a q ) + C 2 H 3 2 ~ ( fl(? ) (Equation 4) 

The equilibrium constant, K a , for acetic acid is 1.8 x 10 _o . Again, this value is small indicating that the 
equilibrium position lies more to the left than to the right. In other words, there are more acetic acid 
molecules at equilibrium than there are acetate ions or hydronium ions. Since K a is less than 1, we know 
that [HC 2 H 3 2 ] > [C 2 H 3 2 ~] or [H + ] at equilibrium. 

Sample question 1: Put the following acids in order of decreasing acid strength. Write equilibrium reactions 
for each. Remember K a is the equilibrium constant for the acid ionization which increases with increasing 
acid strength (or decreasing pH). 

Formic acid, HCOOH, K a = 6.3 x 10~ 4 

Phosphoric acid, H 3 P0 4 ,K a = 7.2 x 10~ 3 

Oxalic Acid, HCOOCOOH,K a = 5.6 x 10~ 2 

Arsenic acid, H 3 AsOi,K a = 6.0 x 10~ 3 . 

Solution: 

Order of decreasing acid strength: Oxalic acid > Phosphoric acid > Arsenic acid > Formic acid 

Equilibrium Reactions: 

1. Oxalic acid: HCOOCOOH {aq) ±? H + {aq) + HCOOCOO {aq) 

2. Phosphoric acid: H 3 PO A ^ aq ) i? H + {aq) + H 2 PO/C \ aq) 

3. Arsenic acid: H 3 AsO i{aq) i? H + \ aq) + H 2 As04T{ aq ) 

4. Formic acid: HCOOH (aq) ±* H + {aq) + COOH \ aq) 

511 www.ckl2.org 



Ka and Kb 

The pH of solutions of strong acids and strong bases can be calculated simply by knowing the concentration 
of the strong acid and, of course, knowing that these acids and bases dissociate in solution 100%. Consider 
a solution that is 0.010 M HC1. HCl is a strong acid and therefore, the acid molecules dissociate 100% and 
this 0.010 M solution of HCl will be 0.010 M in H + ion and 0.010 M in Ct ion. Plugging the value of the 
hydrogen ion concentration into the pH formula, we can determine that this solution has a pH = 2. 

Consider a solution that is 0.0010 M NaOH. NaOH is a strong base and therefore, this solution will be 
0.0010 M in sodium ions and also in hydroxide ions. Since the solution is 1.0 X 10 -3 M in hydroxide ions, 
it will be 1.0 X 10~ n M in hydrogen ions. Therefore, this solution will have a pH = 11. 

A diatomic strong acid such as H 2 S0 4 is only slightly more complicated. Suppose we wish to determine 
the pH of a 0.00010 M solution of H 2 S0 4 . 

H 2 S0 4(aq) -»2 H + + S0 4 2 ~ 

Since the original solution was 0.00010 M in the acid molecules and sulfuric acid is a strong acid, then the 
100% dissociation will produce a solution that contains [H + ] = 0.00020 M. Substituting this hydrogen ion 
concentration in the pH formula yields: 

pH = -log (2.0 x 10~ 4 ) 

pH = -(0.30 - 4) = -(-3.7) = 3.7 

Let's now consider the process for finding the pH of weak acids and bases. In these cases, you need more 
information than you need for strong acids and bases. Not only do you need to know the concentration of 
the original acid or base solution, but you also must know the K a or K/,. 

Suppose we wish to know the pH of a 1.0 M solution of ascorbic acid, H2C%H^O^i a ^\, whose K a = 1.8 x 10 -5 . 

^C^H^O^a^ £5 H ( fl? ) + HCqHqOq {aq) K a = 1.8 X 10 

The K a for this reaction would be written 

[H+][HC 6 H 6 6 -] 



K a 



[H 2 C q H & Oq] 



To find the hydrogen ion concentration from this K a expression and the original concentration of the acid, 
we need a little algebra. We need to assign some variables. 

Let the molarity of the acid that dissociated be represented by x. Then the molarity of the hydrogen 
ions and ascorbate ions in solution will also be represented by x. The molarity of the undissociated acid 
remaining would be represented by 1.0 — x. We can now substitute these variables into the K a expression 
and set it equal to the given K a value. 

(X)(X) r 

K a = , - \ = 1.8 X 10" 5 Expression 1 

(1.0 - x) 

When this equation is simplified, we find that it is a quadratic equation . . . which, of course, can be solved 
by the quadratic formula. 

x 2 + (1.8xl0~ 5 )x-1.8xl0~ 5 = and x = 4.2 x 10~ 3 M 

www.ckl2.org 512 



However, there is a shortcut available to solve this problem that simplifies the math greatly. It involves 
significant figures and adding or subtracting extremely small numbers from large numbers. If you are 
working to 3 significant figures and you are required to subtract 0.00005 from 1.00, when you carry out the 
subtraction and round to 3 significant figures, you discover that you get the original number before you 
subtracted. 

1.00 - 0.00005 = 0.99995 which to 3 significant figures is 1.00 

In the problem we solved above about ascorbic acid, the K a value is very small, 1.8 x 10~ 5 . This indicates 
that the amount of ascorbic acid that dissociates, represented by x, is tiny. When we assigned the variables 
in that problem, we see that the molarity of ascorbic acid remaining after dissociation is represented by 
1.0 - x. Since this x is very tiny, the result of this subtraction will be 1.0 M. Therefore, the K a expression 
from above, 

K a = / A '— = 1.8 X 10~ 5 Expression 1 

(1.0 - x) 

can quite safely be written as 

(x)(x) K „ 

K a = jfftf- = 1-8 x 10~ 5 Expression 2 

You should notice that the variable 1.0 - x has been replaced simply by 1.0. We are assuming that the x 
is so small, it will not alter the value 1.0 when we subtract and then round to proper significant figures. 
This is a safe assumption when the value of the K a is very small (less than 1 x 10~ 3 ). 

When we solve this second expression 

(x)(x) K 

VV = 1-8 xlO" 5 
(1.0) 

x 2 = 1.8xl0~ 5 and 

x = 4.2 x 10~ 3 M 

We no longer need to use the quadratic formula and you must note that the answer to the proper number 
of significant figures is exactly the same as when Expression 1 was solved by the quadratic formula. 

Let's go through another of these problems finding the pH of a weak acid given the initial concentration 
of the acid and its K a . This time we will use a hypothetical weak acid, 0.10 M HA, whose K a = 4.0 x 10 -7 . 

HA (aq) ±+H + + A~ K a = 4.0 x 10~ 7 

[H+M-] - 

K —- - --40x1(1"' 

K a - [ha] - 4.U X 1U 

Let x represent the molarity of HA that dissociates, then [H + ] = [A~] = x, and [HA] = 0.10 - x. 

K a = r ; = 4.0 X 10" 7 

[0.10 - x] 

Once again, since the K a value is very small, the x subtracted in the denominator can be neglected and 
the equation becomes 

Ixl Ixl 7 

! " ' -4.0xl0 -7 



[0.10] 

513 www.ckl2.org 



so, x 2 = 4.0 x 1CT 8 and x = 2.0 x 10~ 4 M. 

Therefore, the hydrogen ion concentration in this solution is 2.0 X 10~ 4 M and substituting this value into 
the pH formula yields 

pH = -log (2.0 x 10~ 4 ) = -(0.3 - 4) = -(-3.7) = 3.7 

The same process is used for weak bases. There is one additional step when working with weak bases because 
once the hydroxide ion concentration is determined, you must then find the hydrogen ion concentration 
before substituting into the pH formula. Kj, represents the equilibrium constant for the dissociation of a 
weak base. Look at the dissociation of dimethylamine (a weak base used in making detergents). We will 
calculate the pH of a 1.0 M solution of this weak base. 

(CH 3 ) 2 NH {aq) + H 2 {L) fc> (CH 3 ) 2 NH 2 + {aq) + OH~ {aq) K b = 5.9 x 10" 4 

WW 7 

[(CH 3 ) 2 NH] 

Allowing x to represent the molarity of {CH 3 ) 2 NH that dissociates results in [{CH 3 ) 2 NH 2 + \ = x and 
[OH—] = x. The molarity of undissociated (CH 3 ) 2 NH will be 1.0 - x. 

Substituting the variables into the Kb expression yields 

Ijcl be] 7 

1 Jl J -4.0xl0~ 7 



[1.0 -x] 



and neglecting the x subtracted in the denominator because it is beyond the significant figures of the 
problem yields 

\x] be] 7 

! ' ! ' - 4.0 x 10" 7 . 



[1.0] 

Therefore, x 2 = 4.0 x 10" 7 and i = 6.3x 10~ 4 M. 

Now that we know the hydroxide ion concentration in the solution, we calculate the hydrogen ion concen- 
tration by dividing the [OH~] into the K w . This will yield [H + ] = 1.6 x 10~ n M. The final step is to plug 
the hydrogen ion concentration into the pH formula. 

pH = -log (1.6 x lO" 11 ) = 10.8 



Sample question 2: Acetic acid is mixed with water to form a 0.10 mol/L HC 2 H 3 2 t aq \ solution at 25° C. If 
the equilibrium concentrations of H 3 + r aq \ and C 2 H 3 2 ~ i aq \ are both 1.34xl0~ 3 mol/L and the equilibrium 
concentration of HC 2 H 3 2 ^ is 0.0999 mol/L, determine the 

(a) K a and 

(b) the pH of the solution at equilibrium. 
Solution: 

HC 2 H 3 2 ( aq ) <=> # + ( a? ) + C 2 H 3 2 ( a? ) 

( a \ K ._ [H+][C 2 H 3 Q 2 -] _ (1.34xl0- 3 )(1.34xl0- 3 ) _ i on v i n-5 
W ^" - IHC 2 H 3 2 ] ~ (0.0999) ~ 1 " OU A 1U 

(b) 

pH = -log [H + ] 

pH = -log (1.34 x 10" 3 ) 

pH = 2.87 



www.ckl2.org 514 



Lesson Summary 

• Weak acids only partially ionize in solution and therefore represent equilibrium reactions. Weak bases 
only partially dissociate in solution and therefore represent equilibrium reactions. K a represents the 
equilibrium constant for the ionization of a weak acid. 

• Kb represents the equilibrium constant for the dissociation of a weak base. Equilibrium calculations 
are the same for weak acids and bases as they were for all other equilibrium reactions. 

Review Questions 

1. What makes weak acids and bases a special case for equilibrium reactions? 

2. What do the constants K a and Kb represent? 

3. Oxalic acid is a weak acid. Its ionization reaction is represented below. H2C2O Uiaq) + H 2 On\ ^> 
H%0 + t aq \ +HC 2 0&~ Uq) Which of the following best represents the acid ionization constant expression, 
K f 

"■a ■ 



(a) K a 

(b) K a 

(c) K a 

(d) K a 



lH 3 0+][HC 2 0- 4 ] 
[H 2 C 2 4 ][H 2 0] 
[H 2 C 2 Q 4 }[H 2 0] 
[H 3 0+][HC 2 0- 4 ] 
[H 3 Q+][HC 2 Q- 4 ] 
W2C2O4] 
[H 2 C 2 Q 4 ] 



[H 3 0+][HC 2 0~ 4 ] 

4. Choose the weakest acid from the list below. 

(a) HN0 2{aq) ;K a = 5.6 X 10" 3 

(b) HF {aq) ;K a = 6.6xlO- 4 

(c) H 3 PO <aq)] K a = 6.9xl(T 3 

(d) HCOOH {aq) ;K a = 1.8 X 1CT 4 

5. Choose one of the following reactions that would best represent a reaction that has an equilibrium 
constant best described as a base dissociation constant, Kb. 

(a) H 2 P0 4 ~ {aq) + H 2 {L) ±5 HP0 4 -\ aq) + H 3 0+ {aq) 

(b) NH A + {aq) + H 2 {L) U NH Haq) + H 3 0+ {aq) 

(c) NH 4 - {aq) + OH- {aq) £+ NH 3{aq) + H 2 {L) 

(d) F- {aq) + H 2 {L) <^ HF {aq) + OH- {aq) 

6. A 0.150 mol/L solution of a weak acid having the general formula HA is 15.0% ionized in aqueous 
solution. Which expression best represents the calculation of the acid ionization constant K a for this 
acid? 

/ s K _ (0.150)(0.150) 

W ^ a ~ (0.150) 

/, n „ _ (0.0225)(0.0225) 

[0) &a - ( 0128 ) 

(c) Not enough information is given. 

7. Put the following bases in order of increasing base strength. Write equilibrium reactions for each, 
ethanolamine (HOCH 2 CH 2 NH 2 ),K h = 3.2 X 10~ 5 , piperidine (C 5 H w NH),K h = 1.3 X 10~ 3 , triethy- 
lamine {{CH 3 CH 2 ) 3 N),K b = 5.2 x 10~ 4 , and ethylenediamine (H 2 NCH 2 CH 2 NH 2 ),K h = 8.5 x 10~ 5 . 

Further Reading / Supplemental Links 

• http://en.wikipedia.org/wiki 

515 www.ckl2.org 



Vocabulary 

acid ionization constant K a represents the equilibrium constant for the ionization of a weak acid. 
base dissociation constant Kb represents the equilibrium constant for the dissociation of a weak base. 

17.7 Br0nsted Lowry Acids-Bases 

Lesson Objectives 

• Define a Br0nsted-Lowry acid and base. 

• Identify Br0nsted-Lowry acids and bases from balanced chemical equations. 

• Define conjugate acid and conjugate base. 

• Identify conjugate acids-bases in balanced chemical equations. 

• Identify the strength of the conjugate acids and bases from strengths of the acids and bases. 

Introduction 

Arrhenius provided chemistry with the first definition of acids and bases but like a lot of theories, this 
model tended to be refined over time. This is exactly what happened in the area of acid/base chemistry. 
Two chemists, named Br0nsted and Lowry, working on similar experiments as Arrhenius, derived a more 
generalized definition for acids and bases that we use in conjunction with the Arrhenius theory. The 
Br0nsted-Lowry theory is the focus of this lesson. As the Br0nsted-Lowry definition unfolded, the number 
of acids and bases that were able to fit into each category increased. Thus, the definitions were broader 
for each. 

Br0nsted-Lowry Acids and Bases 
Br0nsted-Lowry Definitions 

Arrhenius made great in-roads into the understanding of acids and bases and how they behaved in chemical 
reactions. Br0nsted and Lowry slightly altered the Arrhenius definition and greatly enlarged the number 
of compounds that qualify as bases. The Br0nsted-Lowry theory defines an acid as a substance that is a 
proton donor and a base as a proton accepter. Look at equation 1 in which hydrochloric acid is reacting 
with water: 

HCl (s) + H 2°(L) -> H 3 + ^ m]) + Cr^ aq) (Equation 1) 

What is happening is that the acid, HCl is losing a H + ion to form CV and the H2O is gaining a H + ion 
to form H$0 + . The Br0nsted-Lowry concept of acids and bases states that the acid donates a proton 
and the base accepts a proton. Therefore HCl is donating an H + ion to H2O to form Cl~ and the H2O is 
accepting an H + ion from HCl to form H%0 + ; HCl is acting as the acid and H20{aq) is acting as the base. 
Look at equation 2 to see another example of the Bronsted-Lowry Theory in action. 

H 2 POf {aq) + OH- (aq) U HPOi 2 - {aq) + H 2 (L) (Equation 2) 

We can see that again, the equation shows H2PO4 is donating a proton to OH to form HPO^~ and OH 
is accepting the proton to form H2O. Thus H2PO^{aq) is acting as the acid and OH~{aq) is acting as the 
base. 

www.ckl2.org 516 



Sample question: Identify the Br0nsted Lowry Acids and Bases from each of the following equations: 

(a) HC 2 H 3 2 ^ + H 2 0( L ) <=> C 2 H 3 2 ^ + H 3 + ( aq ) 

(b) HCN {aq) + H 2 {L) £5 CN- {aq) + H 3 0+ {aq) 
Solution: 



HC 2 H 3 2 ( aq) + #2<?(L) ±+ C 2 H 3 2 ( a? ) + // 3 ( aq ) 

Acid base 



(b) 



#CAr (a<?) + H 2 (L) i? CN- (aq) + H 3 + {aq) 
Acid base 



Br0nsted— Lowry Acids/Bases Definitions Includes More Compounds 

If you think about the definition of an Arrhenius acid, it includes substances such as HCl, HN0 3 , HC 2 H 3 2 , 
in essence all substances that contain H + ions. This is because according to Arrhenius, the acid ionizes in 
water to produce H + ions. This definition limits what can fit under the umbrella of the definition of acid. 
The Br0nsted-Lowry definition of the acid is broader in that it defines the acid as a proton donor. With 
this broader definition there is the ability to include more compounds in the category of acid. Consider 
the reaction Equation 3: 

HSO A ~ {aq) + OH~( aq) i? H 2 {L) +S0 4 2 ~( aq ) (Equation 3) 

It needs to be pointed out that if a substance is an acid in the Arrhenius definition, it will be an acid in 
the Br0nsted-Lowry definition. However, the reverse is not true. Nor, do Br0nsted-Lowry acids now have 
the properties of an Arrhenius acid. The same is true for bases. In equation 3, the hydroxide ion, OH~, is 
both an Arrhenius base and in the reaction a Br0nsted-Lowry base. In other words, the Br0nsted-Lowry 
definition can be viewed as an extension to the Arrhenius definition rather than a replacement for it. 

May or May Not be in Water Solution 

With the Arrhenius theory, one aspect that is consistent, is that water was part of the equation. Arrhenius 
said that an acid must produce H + ions in a water solution. Therefore, according to Arrhenius, the 
following equation would be representative of an Arrhenius acid. 

HCl {g) + H 2 {L) -> H 3 + {aq) + OH- {aq) (Equation 4) 

If you look at all of the equations we have used thus far for Arrhenius, and even the definitions in Arrhenius 
theory, one commonality shows through: water must be in the equation. Here is where the Br0nsted-Lowry 
definition again varies from Arrhenius theory. Look at the equation below. 

NH 3{aq) + NH 3{aq) ±5 NH 4 + {aq) + NH 2 - (aq) (Equation 5) 

When you look at Equation 5, the first NH 3 molecule is accepting a proton to form NH^ and is therefore 
a Br0nsted-Lowry base, the second NH 3 t aq \ molecule is donating a proton to form NH 2 ~ and is therefore a 

51T www.ckl2.org 



Br0nsted-Lowry acid. Ammonia molecules, however, do not donate hydrogen ions in water and therefore, 
do not qualify as Arrhenius acids. The Br0nsted-Lowry theory has provided a broader theory for acid-base 
chemistry. 

It should be noted that NH 3 is an example of an amphoteric species. Amphoteric species are those that 
in different situations can act as either an acid or a base. That is, in some circumstances, they donate 
a proton and in other circumstances, they accept a proton. Here in Equation 5, NH 3 is doing just this. 
Water is also an amphoteric species. 

Demonstrations explain pH and how it is measured, and the important role of acids and bases. The Proton 
in Chemistry (http : //www . learner . org/vod/vod_window . html?pid=808) 

Acid-Base Conjugate Pairs Definition 

There is one more aspect of the Br0nsted-Lowry theory that was a significant breakthrough to acid-base 
chemistry. Br0nsted and Lowry said that in acid-base reactions, there are actually pairs of acids and bases 
in the reaction itself. According to Bronsted-Lowry, for every acid there is a conjugate base associated 
with that acid. The conjugate base is the result of the acid losing (or donating) a proton. Therefore, if 
you look below, you can see on the left, the acid and on the right the conjugate base. 

Acid Conjugate Base 

HCI -> CI 

HBr -> Br 

HN0 3 -> N0 3 

HC 2 H 3 2 -> C 2 H 3 0~ 2 

Notice that the difference between the acid and its conjugate base is simply a difference of a proton. The 
conjugate base has one less proton (or H + ). Conjugate acids work the same way. For every base in the 
acid-base reaction, there must be a corresponding conjugate acid. The conjugate acid is the result of 
the base gaining (or accepting) the proton. Look below to see the difference between the base and the 
corresponding conjugate acid. 

Base Conjugate Acid 

NH 3 -> NH+ 

OH -> HOH 

H 2 -> H 3 + 

COf -> HCOl 

Now that we know what a conjugate acid and base is, let's now try to identify them in acid-base reactions. 
Look at Equation 6 below, a reaction between acetic acid and water. 

C 2 H 3 2 ~( aq ) + H 3 + {aq) (Equation 6) 





HC 2 H 3 2 (aq) 


+ 


H 2 0( L ) - 




Acetic Acid 




Water 


Step 


1: Identify the acid and base. 




HC 2 H 3 2 ( aq ) 




+ H 2 {L) 




Acid 




Base 



<=> C 2 H 3 2 ( ag ) + H 3 ( aq) 

www.ckl2.org 518 



Step 2: Identify the conjugate acid and base on the product side of the equation. Look at the product side 
to see what product has gained a proton (this is the conjugate base) and which product has lost a proton 
(this is the conjugate acid). 

HC2H- i 2 ( a q) + #2<2(L) ^> C2H3O2 ( a q) + H 3 ( m/ ) 

Acid Base Conjugate Base Conjugate Acid 

As a result, the conjugate acid/base pairs are HC<iHz02{aq) j CiHj,0<i~ (aq) and H2O Vn / H 3 + (aq)- 
Now you try. Identify the conjugate acid-base pairs in the following equation. 

CH 3 NH 2{aq) + HCIO {aq) £5 ClO- {aq) + CH 3 NH 3 + {aq) (Equation 7) 

First, identify the acid and base on the reactant side. 
Step 1: 

CH 3 NH 2{aq) + HClO {aq) i? ClO~ (aq) + CH 3 NH 3 + {aq) 

Base Acid 

Then identify the conjugate acid and base in the products. 
Step 2: 

CH 3 NH 2{aq) + HClO {aq) t; ClO~ {aq) + CH 3 NH 3 + (aq) 

Base Acid Conjugate Base Conjugate Acid 

Hence, the conjugate acid/base pairs are CH 3 NH2/CH 3 NH^ and HCIO / ClO~ . 

Sample Problem: Identify the conjugate acid-base conjugate pairs in each of the following equations: 

(a) NH 3{aq) + HCN {aq) ±+ NH 4 + {aq) + CN~ (aq) 

(b) C0 3 2 - {aq) + H 2 {L) U HC0 3 - {aq) + OH- {aq) 
Solution: 

(a)NH 3 /NH+ and HCN/CN- 
(b)C0 3 2 -/HC0 3 - and H 2 0/OH- 

The Strength of Conjugate Acids and Bases 

The definition of a Br0nsted-Lowry acid is that it donates a proton (hydrogen ion) . In order to be a strong 
Br0nsted-Lowry acid, it must donate the proton readily. That is, the bond between the hydrogen ion 
and the cation is weak so that the proton is released easily. The definition of a Bronsted-Lowry base is 
that it accepts a proton. In order to be a strong base, it must take on that proton readily. Therefore, a 
Br0nsted-Lowry strong base would hold onto the proton tightly. 

If we consider a strong acid, such as HCl, we know that the acid releases its proton readily . . . does that 
tell us anything about the Cl~ ion as a base? If the chloride ion does not hold onto the proton tightly, we 
know HCl will be a strong acid, and therefore, because the chloride ion releases the proton easily, it cannot 
be a strong base. The same behavior that qualifies an anion attached to a hydrogen ion as a strong acid 
also qualifies the anion alone as a weak base. 

519 www.ckl2.org 



The reverse of this situation would also be true. Consider a weak acid such as acetic acid, HC2H%02- This 
is a weak acid because the acetate ion does not release the hydrogen ion readily; it holds it tightly, hence 
only ionizes to a slight degree. That tells us that the acetate ion takes on a proton readily and holds it 
and that qualifies the acetate ion as a strong base. 

Strong acids produce conjugate bases that are weak bases and weak acids produce conjugate 
bases that are strong bases. 

In this same way, a strong base such as OH~ ion takes on hydrogen ions readily and holds them tightly. 
Therefore, the conjugate base of OH~ ion, which is H2O, will be a weak acid. A weak base such as ammonia, 
NH3, does not take on a hydrogen ion readily and that causes it to be a weak base. Therefore, the conjugate 
acid of NH3, NH4 + , will be a strong acid. 

Strong bases produce conjugate acids that are weak acids and weak bases produce conjugate 
acids that are strong acids. 

See Table 17.8, which conjugate bases would be strong and which would be weak? 

Table 17.8: Strong Acids 



Name 



Symbol 



Conjugate Base 


Strength of Conjugate 




Base 


cr 


Weak 


Br 


Weak 


I 


Weak 


N0 3 - 


Weak 


ClO A - 


Weak 


HSO4- 


Weak 



Hydrochloric Acid 


HCl 


Hydrobromic Acid 


HBr 


Hydroiodic Acid 


HI 


Nitric Acid 


HN0 3 


Perchloric Acid 


HClO A 


Sulfuric Acid 


H2SO4 



All of the strong acids will have weak conjugate bases. All other acids are weak and will therefore have 
conjugate bases that are strong. 

The stronger the acid, the weaker the conjugate base. This means that the conjugate bases will follow a 
similar trend as the acids only in the reverse. Look at Table 2. See how the acid strength trend compares 
with the conjugate base strength trend. 

Table 2: Strong Acids and Their Conjugate Bases 



Increasing 

Acid 
Strength 



Strong Acid 


Conjugate Base 


HCl 


CI" 


HBr 


Br" 


HI 


1" 


HNOj 


NOj 


HCIO a 


cio; 


H z S0 4 


hso; 



Increasing 

Base 
Strength 



As we have seen with the strong acids-weak conjugate bases and weak acids-strong conjugate bases, the 
same concept can be applied to bases. Strong bases have weak conjugate acids and weak bases have strong 
conjugate acids. Which conjugate acids would be strong and which would be weak? 

Only the OH— ion represents a strong base and therefore only H2O represents a weak conjugate acid. All 
of the other bases are weak. Therefore all of the other conjugate acids are strong. The strong bases are 
LiOH,NaOH,KOH,RbOH,CsOH,Ca(OH) 2 ,Sr(OH)2, and Ba(OH) 2 - All of these strong bases contain OH~ 
and therefore their weak conjugate acid will be H2O. The remaining bases are weak and so would have 



www.ckl2.org 



520 



strong conjugate acids. 

Lesson Summary 

• The Br0nsted-Lowry concept of acids and bases states that the acid donates a proton and the base 
accepts a proton. 

• The Br0nsted-Lowry definition includes more substances because the definition is not confined to 
substances containing hydroxides or to be in water. A conjugate acid is a substance that results 
when a base gains (or accepts) a proton. 

• A conjugate base is a substance that results when an acid loses (or donates) a proton. 

• Strong acids result in weak conjugate bases when they lose a proton and weak acids result in strong 
conjugate bases when they lose a proton. 

• Strong bases result in weak conjugate acids when they gain a proton and weak bases result in strong 
conjugate acids when they gain a proton. 

Review Questions 

1. What improvements did Br0nsted-Lowry make over the Arrhenius definition for acids-bases? 

2. If you were to use HCl{aq) as an example, how would you compare the Arrhenius definition of an 
acid to the Br0nsted-Lowry definition? 

3. What is the Br0nsted-Lowry definition of an acid? 

(a) a substance that donates protons 

(b) a substance that accepts protons 

(c) a substance that dissolves in water to form OH' ions 

(d) a substance that dissolves in water to form H + ions 

4. If H 3 + is an acid according to the Br0nsted-Lowry theory, what is the conjugate base of this acid? 

(a) H 4 2+ {aq) 

(b) H+ {aq) 

(c) H 2 {L) 

(d) OH~ {aq) 

5. What is the conjugate base of H 2 P0 4 ^1 

(a) H 3 + {aq) 

(b) H 3 P04 (aq) 

(c) HP0 4 2 - {aq) 

(d) P0 4 3 - {aq) 

6. In the following reactions, which are the Br0nsted-Lowry acids? i) H 3 PO^ aq ^ + H 2 0^ ^=> HzO + ( aq ) + 
H2PO A - {aq) ii) H 2 P0 4 - {aq) + H 2 {L) £> H 3 0+ (aq) + HPO A 2 ~ {aq) 

(a) H 2 PO A ~,H20,HPO A 2 ~ 

(b) H 3 P0 4 ,H 2 0,H 2 P0 4 ~ 

(c) H 3 0+,H 2 0,HPO A 2 - 

(d) H 3 P04,H 3 + ,H 2 P0 4 

521 www.cki2.0rg 



7. Label the conjugate acid-base pairs in each reaction. 

(a) HC0 3 + H 2 0<=> H 2 C0 3 + OH 

(b) H 2 POi + H 2 «=> H 3 + + HP0 4 2 ~ 

(c) cr + Zf 2 t? //CN + OH 

(d) #F(a ? ) + // 2 0(/) fl H 3 + {aq) + F-(aq) 

8. Complete the following reactions. When done, label the conjugate acid/base pairs. 

(a) Br0 3 +H 2 0^ 

(b) HN0 3 + H 2 -> 

(c) //SOf + C 2 Oi 2 ~ <=> 

9. For the reactions in question 7, which are the weak conjugate bases and which are the strong conjugate 
bases? 

Further Reading / Supplemental Links 

• http : //learner . org/resources/series61 . html 

The learner.org website allows users to view streaming videos of the Annenberg series of chemistry videos. 
You are required to register before you can watch the videos but there is no charge. The website has one 
video that relates to this lesson called The Proton in Chemistry. 

Vocabulary 

Br0nsted-Lowry acid A substance that donates a proton (H + ). 

Br0nsted-Lowry base A substance that accepts a proton (H + ). 

amphoteric substances Substances that act as both acids and bases in reactions (i.e. NH 3 ). 

conjugate acid The substance that results when a base gains (or accepts) a proton. 

conjugate base The substance that results when an acid loses (or donates) a proton. 

17.8 Lewis Acids and Bases 

Lesson Objectives 

• Define a Lewis acid and a Lewis base. 

• Understand and define a coordinate covalent bond. 

• Identify a Lewis acid and a base in reactions. 

Introduction 

Lewis is known for the shared-pair chemical bond and his work with electrons. We will be concentrating on 
his work with electrons. You may recall working with Lewis electron dot diagrams from an earlier chapter. 
Look at the example below for a quick refresher. 

www.ckl2.org 522 



CH 4 H:c:H 



H 
C 

h 



In the early 1930s, Gilbert Lewis saw the need for a more general definition for acids and bases that 
involved the use of electrons. Using the work of Br0nsted and Lowry, he saw that in some cases the 
acids may not have protons to donate, but may have electron pairs to donate. We will give a look at the 
contribution of Lewis to the theory of acids and bases. It should be noted that most chemists today use the 
Br0nsted-Lowry definitions of acids-bases and the Lewis definitions are used in more specialized situations. 



Lewis Acids 

Lewis defined an acid as a substance that accepts a pair of electrons from another substance. Therefore, 
Lewis acids must have room in their structure to accept a pair of electrons. Remember that each central 
atom can hold eight electrons. What this means is that if the atom does not have eight electrons, but six 
electrons, then it can accept one more pair. Look at the structure in Figure 17.5. Notice how, in each 
case, there is room to accept a pair of electrons. 

The Lewis acid will accept the electron pair in order to form a bond. The bond that forms between the two 
atoms will be covalent bonds. Covalent bonds are formed when electrons are shared between two atoms. 




H + 



The BF 2 molecule can hold This hydrogen ion can hold 2 

3 electrons surrounding the electrons around the hydrogen 

boron atom ... but in this atom ... but in this case has 

case has only 6 electrons zero electrons around the 

around the boron. Therefore. hydrogen. Therefore, this ion 

this molecule can take on can take on two electrons and 

two electrons and that makes that makes it a Lewis acid, 
it a Lewis acid. 



Figure 17.5: Examples of Lewis Acids. 



Lewis Bases 

Lewis defined a base as a substance that donates a pair of electrons to a substance. Lewis bases have a 
lone pair of electrons that they can share with another species when they donate them as part of their 
performance as a base. If we look at the example of bases in Figure 17.6, we can see that each of them 
has a lone pair of electrons available to donate. 

Again, since the base will share the electrons with the corresponding acid when it donates this pair, the 
resulting bond that forms between the acid and the base will be a covalent bond. 

523 www.ckl2.org 



H 
H:n: H: ° ; 



ii\ 



This ammonia molecule has This hydroxide ion has four 

four pairs of electrons around pairs of electrons around its 

its central atom and only three central atom and only one 

of the pairs are shared. of the pairs is shared. 

Therefore, this molecule has Therefore, this ion has unshared 

another pair of electrons that pairs of electrons available to 

it can share and that makes share and that makes it a 

it a Lewis base. Lewis base. 



Figure 17.6: Examples of Lewis Bases. 



Lewis Acid-Base Neutralization 

We have said already that the Lewis acid accepts a pair of electrons from a Lewis base that donates the 
pair of electrons. The resulting bond that forms is a covalent bond. In a regular covalent bond, each of 
the bonded atoms contributes one of the shared pair of electrons. In certain cases, it is possible for one of 
the bonded atoms to contribute both of the shared electrons and the other atom contributes no electrons. 
Coordinate covalent bond is the term given to bonds formed when both electrons come from the same 
atom. Look at the equation below. 

H IF H F 

H : N : + B : F ^- H : N : B : F 

H F H F 

This is a Lewis acid combining with a Lewis base and therefore, this reaction represents a Lewis acid-base 
neutralization reaction. Since both of the electrons that participate in the bond between TV and B have 
come from the Lewis base NH 3 , the N B bond is a coordinate covalent bond. 

Sample question: Identify the Lewis acid and Lewis base in each of the following reactions. Then write 
the reactions to show that these reactions involve coordinate covalent bonds. 

(a) Ag + {aq) + 2NH 3{ay) -» Ag{NH 3 ) 2(aq) 

(b) H 2 {L) + NH 3{aq) -> NH 4 OH {aq) 
Solution: 

(a)Ag+ K) + 2NH 3[aq] -> Ag{NH z ) 2{aq) 
Acid Base 

H H H H 

h : n : + Ag + : n : h — *■ h : n : Ag : n : h 

H H H H 

(a)tt 2 (L) + NH 3{aq) -> NH 4 OH {aq) 
Acid Base 

www.ckl2.org 524 



H H 

■ ■ ■■ ■ ■ ■■ 

h : n : + h:o: — > h : n :h:o: 

■■ ■■ ■■ ■■ 

H H H H 



Lesson Summary 

• Lewis defined an acid as a substance that accepts a pair of eiectrons from a substance to form a 
bond. Lewis denned a base as a substance that donates a pair of electrons to a substance to form a 
bond. 

• Coordinate covalent bond is the term given to bonds formed when both electrons come from the 
same atom. 

Review Questions 

1. How do the Lewis definitions of acids and bases compare to the Br0nsted-Lowry definitions of acids 
and bases? 

2. In the following reversible reaction, which of the reactants is acting as a Lewis base? Cd 2+ i aq \ + 



Ar {aq) u cdh 2 - 


(aq) 












(a) 


Cd 2 + 














(b) 


I 














(c) 


Cdh 2 ~ 














(d) 


none of the above, 


this 


is 


not 


an 


acid-base reaction 



3. Which of the following statements are false? 

(a) NH 3 is a Lewis base. 

(b) B{OH) 3 is a Lewis acid. 

(c) CO2 is a Lewis base. 

(d) Ag + Is a Lewis acid. 

4. Which of the following statements are true? 

(a) NH 3 is a Lewis base. 

(b) B(OH) 3 is a Lewis acid. 

(c) CO2 is a Lewis base. 

(d) Ag + Is a Lewis acid. 

5. Classify each of the following as a Lewis acid or base. 

(a) H 2 

(b) BF 3 

(c) S 2 - 

(d) Cu 2+ 

(e) O 2 - 

6. Write the balanced chemical equation between S0 3 2 ~ and H2O and label the Lewis acids and bases. 

7. Identify the Lewis acid and Lewis base in each of the following reactions. Then write the reactions. 

(a) Cu 2+ (aq) + 6 H 2 {L) -* Cu(H 2 0) 6 2 + {aq) 

(b) (CH 3 CH 2 ) 2 (aq) +AlCl Haq) -» (CH 3 CH 2 )20AlCl 3{aq) 

525 www.cki2.0rg 



Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

Lewis acid A substance that accepts a pair of electrons from a substance (i.e. BF3). 

Lewis base A substance that donates a pair of electrons from a substance (i.e. NH3). 

coordinate covalent bond A covalent bond formed where both electrons that are being shared come 
from the same atom. 



Image Sources 



(1) Examples of Lewis Acids.. 

(2) Theresa Forsythe. The pH Scale.. CC-BY-S A. 

(3) pH Scale for Common Substances.. 

(4) Examples of Lewis Bases.. 

(5) Theresa Forsythe. . CC-BY-SA. 

(6) Richard Parsons. Hydrated ions in solution.. CC-BY-SA. 



www.ckl2.org 526 



Chapter 18 

Water, pH and Titration 



18.1 Water Ionizes 

Lesson Objectives 

The student will write the equation for the autoionization of water and express the concentration of 
hydrogen and hydroxide ion in a neutral solution at 25° C. 

The student will express the value of K w in a water solution at 25°C. 

The student will write the formulas for pH and pOH and show the relationship between these values 
and K w . 

The student will express the relationship that exists between pH,pOH, and K w . 

Given the value of any one of the following values in a water solution at 25°C, the student will 
calculate all the other values; [H + ], [OH + ], pH, and pOH. 

The student will state the range of values for pH that indicate a water solution at 25°C is acidic. 

The student will state the range of values for pH that indicate a water solution at 25°C is basic. 

The student will state the range of values for pH that indicate a water solution at 25° C is neutral. 

Introduction 

There are many properties of water that we have learned. We know, for example, that water is called the 
universal solvent because of its ability to dissolve both polar substances as well as most ionic ones. We also 
know that pure water does not conduct electricity. The reason pure water does not conduct electricity is 
because of the small concentration of ions present when water ionizes. In this lesson, we will look a little 
closer at the uniqueness of autoionization. We will expand on the effect other added substances have on 
the [//3<9 + ] and [OH~] in water. Thus, we need to have formulas readily available to calculate the pH, the 
pOH, the [H^O + ], and/or the [OH~]. Much of what you will learn may be review. 

527 www.ckl2.org 



Water Ionization is Small But Significant 

Autoionization, as we learned in a previous lesson, is the process where the same molecule acts as both 
an acid and a base. Water is one of the compounds that can exhibit this unique property. When speaking 
about the autoionization of water, another term that is also used is the self-ionization of water. The 
reaction for the autoionization of water is shown in Equation 1 and its net ionic equation in Equation 2. 

H 2 [L) + H 2 {L) £» H 3 + iaq) + OH- {aq) (Equation 1) 

H + + OH + H 2 0±+ H 3 + (ag) + OH- (aq) (Equation 2) 

If you notice in Equation 1, one H 2 Ou\ molecule donates a proton and is therefore a Br0nsted-Lowry acid; 
the second H 2 On\ molecule accepts a proton and is therefore a Br0nsted-Lowry base. 

While this does indeed occur, it only happens to about 2 molecules for every billion! This is the reason 
that pure - or what can also be referred to as distilled or deionized water - will not conduct electricity. In 
contrast, water that comes into our houses and we get from turning on the tap, is not pure water and can 
conduct electricity. You may have seen hair dryers with warnings labels attached that remind you not to 
use them while in the bath. Tap water has various ions dissolved in it, which makes it an electrolyte and 
an electrical current that could pass through tap water and into you, could be dangerous. Some of the ions 
that are dissolved in tap water are Na + , Cl~, Mg 2+ , Ca 2+ and various others. When tap water boils away 
it often leaves a residue in the pan or kettle. 

The Mathematics of pH and pOH 

Water ionizes to a very slight degree. 

H 2 {L) <=> H + + OH 

At 25°C, the ionization equation above reaches equilibrium when [H + ] = [OH~] = 1 X 10 M. Therefore, 
the equilibrium constant for this reaction, at 25°C, will be: 

K w = [H + ][OH~] = (1 x 1(T 7 )(1 x 1(T 7 ) = 1 x 1(T 14 

When the concentrations of the hydrogen and hydroxide ions are equal, we say the solution is neutral. 
When some substance is added to the water and alters the hydrogen and hydroxide ion concentrations so 
that they are no longer equal, the solution will become either acidic or basic. The hydrogen and hydroxide 
ions will return to an equilibrium position but the concentrations of the ions may not be equal. 

If, at equilibrium, [H + ] > [OH~], the solution is acidic. 
If, at equilibrium, [H + ] = [OH~], the solution is neutral. 
If, at equilibrium, [H + ] < [OH~], the solution is basic. 

pH is used to express the acid strength of water solutions. 

pH = -log [H + ] 

For acid solutions, the pH will be less than 7. For basic solutions, the pH will be greater than 7. For 
solutions that are neutral, the pH will be equal to 7. 

www.ckl2.org 528 



The pH equation can also be used to find the hydrogen ion concentration when the pH is given. 

[H+] = l(T p// 

Even though both acidic and basic solutions can be expressed by pH, someone decided to create an 
equivalent set of expressions for the concentration of the hydroxide ion in water; hence pOH. 

pOH = -log [0H~] 

If the pOH is greater than 7, the solution is acidic. If the pOH is equal to 7, the solution is neutral. If the 
pOH is less than 7, the solution is basic. 

If we take the negative log of the complete K w expression: 

K w =[H+][OH-} 
-log K w = (-log [H+]) + (-log [OH-]) 
-log (1 x 1(T 14 ) = (-log [//+]) + (-log [OH-]) 
14 = pH + pOH. 

We can determine that the sum of the pH and the pOH is always equal to 14 (at 25°C). 

Remember that the pH scale is written with values from to 14 because many useful acid and base solutions 
fall within this range. We can use this formula to carry many different types of calculations. Before moving 
on to these, one more deduction can be made from all that we have done so far. If p() = — log() such that 
pH = -\og[H + ] and pOH = —\og[OH~] and we know that K w = 1.00 x 10~ 14 , then we can calculate pK w 
using the formula pK w = -log K w . 

Now let's go through a few examples to see how this calculation works for problem-solving in solutions 
other than water. 

Sample question 1: What is the [H + ] for a solution of NH 3 whose [OH~] = 8.23 X 10~ 6 mol/L? 

Solution: 

[H 3 + ][OH-] = 1.00 xl0~ 14 

.. 1.00 X10 -14 1.00 xlO -14 q 

[H 3 + ] = = = = = = 1.26 x 10~ 9 M 

L d J [OH-] [8.23X10- 6 ] 

Sample question 2: Black coffee has a [H 3 + ] = 1.26 x 10~ 5 mol/L. What is the pOHl 
Solution: 

pH = -log [H + ] = -log 1.26 x 10~ 5 = 4.90 
pH + pOH = 14 

pOH =U-pH=U- 4.90 = 9.10 

Lesson Summary 

• Autoionization is the process where the same molecule acts as both an acid and a base: 

• pK w = -logK w 

. pH + pOH = pK w = 14.0 

• pH + pOH = 14.0 is a formula used to find a number of different pieces of information from an acid- 
base equation including [H + ], [OH~], pH, and pOH. Only one of these four pieces must be known to 
find the other three. 

529 www.ckl2.org 



Review Questions 

1. How do pH and pOH relate to the pH scale? 

2. What does the value of K w tell you about the autoionization of water? 

3. If the pH of an unknown solution is 4.25, what is the pOHl 

(a) 1(T 4 - 25 

(b) 1(T 9 - 75 

(c) 9.75 

(d) 14.0 -1(T 9 - 75 

4. A solution contains a hydronium ion concentration of 3.36 X 10~4 mol/L. What is the pH of the 
solution? 

(a) 3.36 

(b) 3.47 

(c) 10.53 

(d) none of the above 

5. A solution contains a hydroxide ion concentration of 6.43 x 10~ 9 mol/L. What is the pH of the 
solution? 

(a) 5.80 

(b) 6.48 

(c) 7.52 

(d) 8.19 

6. An unknown solution was found in the lab. The pH of the solution was tested and found to be 3.98. 
What is the concentration of hydroxide ion in this solution? 

(a) 3.98 mol/L 

(b) 0.67 mol/L 

(c) 1.05 x 10~ 4 mol/L 

(d) 9.55 x 10' 11 mol/L 

7. If a solution is known to have a hydroxide ion concentration of 2.5 X 10 -5 mol/L, then the pH of the 
solution is and it is . 

(a) 2.5, acidic 

(b) 4.6, basic 

(c) 4.6, acidic 

(d) 9.4, basic 

8. K w is the ionization product constant for water but is also the equilibrium constant for the acid- 
base autoionization reaction for water. When dealing with equilibrium constants, such as K w , it is 
important to take into account the temperature as temperature affects the value of the equilibrium 
constant. The value of 1.0 x 10~ 14 for K w is for a temperature of 25° C. If the temperature was raised 
to 60°C, the value of K w changes to 1.0 X 10~ 13 . How does this effect [H+],[OH~], pH, and pOHl 

Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

autoionization The process where the same molecule acts as both an acid and a base. 
www.ckl2.org 530 



18.2 Indicators 

Lesson Objectives 

• Define an acid-base indicator. 

• Explain the difference between natural and synthetic indicators. 

• List examples of natural and synthetic indicators. 

• Explain how indicators work. 

• Explain the usefulness of indicators in the lab. 

Introduction 

Indicators are used everyday both in our lives and in our laboratories. We make a cup of tea, we work in 
the garden, we have a bowl of cereal and throw some blueberries on it, or we may go to chemistry class 
and do an experiment to identify some of the properties of household substances using indicator solutions. 
All of these situations involve the use of indicators; the only difference is that some indicators are natural 
and some are synthetic (well, maybe - depending what you use in your chemistry class!) In this lesson you 
are going to examine the concept of natural and synthetic indicators and the use of these in the modern 
chemistry lab. 

Substances That Change Color Due to pH Change 

Litmus paper is a paper that has been dipped in a substance that will undergo a color change when it is 
dipped in either an acid or a base. The litmus paper is called an indicator because it is used to indicate 
whether the solution is an acid or a base. If the red litmus paper turns blue, the solution is basic {pH > 7), 
if the blue litmus turns red the solution is acidic {pH < 7) . 




Figure 18.1: Hydrangeas 

An indicator is a substance that changes color at a specific pH and is used to indicate the pH of the 
solution. The juice from red cabbage, for example, can be used to prepare an indicator paper. It contains 
the chemical anthrocyanin which is the active ingredient in the indicator. Did you know that there are 
actually indicator fish in the world? It's true! In Singapore there are fish that have been genetically altered 
with the indicator in hopes of detecting pollutants in the water. These rainbow fish turn color from green 
to red depending on the amounts of heavy metals and are thus given their name. (You may see, in news 
articles, these fish referred to as "litmus fish" but, of course, the color changes in these fish have nothing to 
do with the substance litmus.) Another example of a natural indicator is flowers. Hydrangea is a common 
garden plant with flowers that come in many colors depending on the pH of the soil. If you are travelling 

531 www.ckl2.org 



around and see a hydrangea plant with blue flowers (Figure 18.1), the soil is acidic, the creamy white 
flowers indicate the soil is neutral, and the pink flowers mean the soil is basic. 

A natural indicator is an indicator that is produced from a substance that is naturally occurring or is 
itself a naturally occurring substance. Red beets, blueberries, and cranberries are other great examples of 
a naturally occurring indicator. These are all due to the same active ingredient anthocyanin found in the 
red cabbage. 

So how do naturally occurring indicators differ from synthetic indicators? Synthetic indicators are 
compounds created in a chemistry lab rather than compounds found in nature. Both naturally occurring 
indicators and synthetic indicators are weak organic acids or bases. For example, a common synthetic 
indicator used in most chemistry laboratories is phenolphthalein (Figure 18.2). 



HO 




OH 



Figure 18.2: Structure of Phenolphthalein. 

This indicator, in particular, changes color at pH of 8.2, before 8.2 it is colorless and after 8.2 it is pink. 
There are many common synthetic indicators that are useful in the chemistry laboratory. All of these 
have similar structures. In the acidic range, chemistry students may use methyl orange. The structure for 
methyl orange is shown in Figure 18.3. 




O Na + 



Figure 18.3: Structure of Methyl Orange (Yellow Form). 

Methyl Orange changes color from pH 3.2 to 4.4. Below 3.2, the color of the indicator is red. Above 4.4 
the color of the indicator is yellow. In between 3.2 and 4.4, at 3.8, the color would be orange ... hence the 
name. 

What happens chemically when there is a color change is that the structure of the molecule donates or 
accepts a proton. Remember Br0nsted-Lowry acids and bases? A Br0nsted-Lowry acid donates a proton 
and a Br0nsted-Lowry base accepts a proton. There are two requirements for a substance to function as 
an acid-base indicator; 1) the substance must have an equilibrium affected by hydrogen ion concentration, 
and 2) the two forms of the compound on opposite sides of the equilibrium must have different colors. Most 



www.ckl2.org 



532 



indicators function in the same general manner and can be presented by a generic indicator equation. In 
the equation below, we represent in the indicator ion with a hydrogen ion attached as HIn and we represent 
the indicator ion without the hydrogen attached as In - . 

The color of an indicator depends on the pH of the solution. (Source: Richard Parsons. CC-BY-SA) 



ffln 



fan) 



«r 



± H+ + 



In" 



Since the indicator itself is a weak acid, the equilibrium between the protonated form and the anionic form 
is controlled by the hydrogen ion concentration. For the example above, the protonated form is colored 
red and the anionic form is colored yellow. If we add hydrogen ion to the solution, the equilibrium will be 
driven toward the reactants and the solution will turn red. If we add base to the solution (reduce hydrogen 
ion concentration), the equilibrium will shift toward the products and the solution will turn yellow. It is 
important to note that if this indicator changes color at pH = 5, then at all pH values less than 5, the 
solution will be red and at all pH values greater than 5, the solution will be yellow. Therefore, putting this 
indicator into a solution and having the solution turn yellow does NOT tell you the pH of the solution . . . 
it only tells you that the pH is greater than 5 ... it could be 6, 7, 8, 9, etc. At pH values less than 5, the 
great majority of the indicator molecules are in the red form and the solution will be red. At pH's greater 
than 5, the great majority of the indicator particles will be in the yellow form and the solution will be 
yellow. The equilibrium between these indicator particles is such that the particles will be 50% red form 
and 50% yellow form at exactly pH = 5. Therefore, at pH = 5, the actual color of the solution will be a 
50 - 50 mixture of red and yellow particles and the solution will be ORANGE. 

The color change of an indicator occurs over a very short range. (Source: Richard Parsons. CC-BY-SA) 



□ □ □ 



pH = 2 



pH = 3 P H = 4 



< < < 



pH = 



P H = 6 



P H = 



pH = 8 



You should recognize that an indicator will only indicate the actual pH at one particular pH and that is 
the color-change pH. 

Laboratory Uses for pH Indicators 

There are many indicators that are available to be used to help determine the pH of solutions. A list 
of the most common indicators is found in Table 18.1 along with their respective color change pHs and 
corresponding color changes. 

Table 18.1: Colors and pH Ranges for Common Indicator 



Indicator 



pH range 



Color Range 



Methyl violet 
Thymol blue 
Orange IV 
Methyl orange 
Bromphenol Blue 
Congo Red 



0.0-1.6 
1.2-2.8 
1.3-3.0 
3.2-4.4 
3.0-4.7 
3.0-5.0 



Yellow - Blue 
Red - Yellow 
Red - Yellow 
Red - Orange 
Orange/Yellow 
Blue - Red 



Violet 



533 



www.ckl2.org 



Table 18.1: (continued) 



Indicator 



pH range 



Color Range 



Bromocresol green 
Methyl red 
Litmus 

Chlorophenol Red 
Bromothymol blue 
Phenol Red 
Thymol blue 
Phenolphthalein 
Thymolphthalein 
Alizarin yellow R 
Methyl Blue 
Indigo Carmine 



3.8 - 5.4 

4.8 - 6.0 
5.0-8.0 
4.8 - 6.2 
6.0 - 7.6 
6.4 - 8.2 
8.0 - 9.6 
8.2 - 10.0 
9.4 - 10.6 
10.1-12.0 
10.6-13.4 
11.4-13.0 



Yellow - Blue 
Red - Yellow 
Red - Blue 
Yellow - Red 
Yellow - Blue 
Yellow -Red/Violet 
Yellow - Blue 
Colorless - Pink 
Colorless - Blue 
Yellow - Red 
Blue - Pale Violet 
Blue - Yellow 



There are many more indicators than are shown in the table above. Those that are shown are ones that 
you may find in common chemistry classroom laboratories or in universities depending on where you are 
located or what your experimental needs are at the time. One that is not found in the table is known as 
the universal indicator. The universal indicator is a solution that has a different color for each pH from 
0-14. Universal indicator is produced by creatively mixing many of the individual indicators together so 
that a different coor is achieved for each different pH. It is used for all types of experiments to determine 
if solutions are acids or bases and where on the pH scale the substance belongs. Figure 18.4 indicates the 
colors of universal indicator for different pH values. 

Color of Universal Indicator at Various pH Values 



1 2 


3 


. 


■ 


• 










1 


8 9 


10 11 


12 13 14 



Figure 18.4: The color of Universal Indicator at various . 

Why is this information be so important? If you look at Table 18.1, Thymol blue changes color from pH 
of 8.0 to pH = 9.6. What this means is that from pH of to pH of 8.0, the color of the solution will stay 
yellow. At pH = 8.0 the color will begin to change from yellow to blue and from pH of 9.6 until pH = 14, 
the color of the solution will be blue. Around pH = 8.8 (the midway mark between 8.0 and 9.6) the color 
of the solution would be green. 

Sample question: If the pH of the solution is 4.8, what would be the color of the solution if the following 
indicators were added? 

(a) Universal indicator 

(b) Bromocresol Green 

(c) Phenol red 
Solution: 

(a) Universal indicator = Orange to orange-yellow (see Figure 18.4) 

(b) Bromocresol Green = green (midway pH = 4.6) 

(c) Phenol red = yellow 



www.ckl2.org 



534 



There is another important use for indicators in the lab and this is to determine the pH of an unknown 
solution. They are also used in titrations, a concept we will be exploring in the another lesson. For now, 
let's look at how we can use indicators to determine the pH of an unknown solution. For example, a 
solution has been found in the laboratory that was tested with a number of indicators. It was found that 
the following indicators showed these results: 

Phenolphthalein was colorless 

Bromocresol green was blue 

Methyl red was yellow 

Phenol red was yellow 

What was the pH of the solution? Let's look at the data. 

Phenolphthalein was colorless, pH < 8.0 

Bromocresol green was blue, pH > 5.4 

Methyl red was orange, pH > 6.2 

Phenol red was yellow, pH < 6.4 

Therefore the pH of the solution must be between 6.3 and 6.4. 

Lesson Summary 

• An indicator is a substance that changes color at a specific pH and is used to indicate the pH of 
the solution. A natural indicator is an indicator that is produced from a substance that is naturally 
occurring or is itself a naturally occurring substance. 

• Synthetic indicators are normally an organic weak acid or base with usually a complicated structure. 
Universal indicator is a solution that has a different color for each pH from to 14. 

Review Questions 

1. Describe the uses for litmus and universal indicator in the laboratory setting. 

2. What is the difference between a natural and a synthetic indicator? 

3. Describe how indicators work. 

4. If you had an acid-base neutralization reaction that turned phenolphthalein pink and Thymolph- 
thalein blue, what is the pH of the solution? 

(a) 8.2 

(b) 9.4 

(c) 10 

(d) Not enough information is available. 

5. If you had an acid-base neutralization reaction that turned methyl violet blue and Thymol blue 
orange, what is the pH of the solution? 

(a) 1.6 

(b) 2.0 

(c) 2.8 

(d) Not enough information is available. 

6. Universal indicator is an indicator commonly used in the laboratory. At a pH of 6 it is pale yellow 
and at a pH of 4 it is pale orange. If the indicator was orange, which statement would be definitely 
true? 

535 www.ckl2.org 



(a) The solution is probably acidic. 

(b) The pH is between 4 and 5. 

(c) The solution is probably basic. 

(d) The pH is less than 5.0. 

7. Alizarin Yellow 7? is an indicator that changes color in the pH range from 10.1 to 12.0. Below 10.1 
the color is Yellow, above 12.0 the color is red. If the color of the solution containing Alizarin Yellow 
R was orange, which statement about the solution would be true? 

(a) The pH is below 10. 

(b) The pH is above 12.0. 

(c) The solution is definitely acidic. 

(d) The pH is between 10.1 and 12.0. 

8. If the pH of the solution is 8.9, what would be the color of the solution if the following indicators 
were added? 

(a) Universal indicator 

(b) Thymol blue 

(c) Methyl blue 

9. A solution has been found in the laboratory that was tested with a number of indicators. It was 
found that the following indicators showed these results: 

(a) Phenolphthalein was colorless 

(b) Orange IV was yellow 

(c) Universal indicator was orange 

(d) Methyl orange was red 

What was the pH of the solution? 

Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

indicator A substance that changes color at a specific pH and is used to indicate the pH of the solution. 

natural indicator An indicator that is produced from a substance that is naturally occurring or is itself 
a naturally occurring substance. 

synthetic indicator An indicator that is a complicated structure of an organic weak acid or base. 

Possible Laboratory Activities for Indicators 
Teacher's Pages for pH Measurements Using Indicators 

Investigation and Experimentation Objectives 




www.ckl2.org 536 



In this activity, the student will observe the relationships between theoretical mathematical conclusions 
and laboratory results. 

Lab Notes 

Buffered solutions of various pH values can be purchased in dropper bottles, as can dropper bottles of 
indicator solutions. If you have many chemistry classes and perform the experiment for several years, it 
may be more economical to prepare the solutions yourself. 

The solutions used in this lab can be prepared as follows. 

pH = 1 solution: dilute 8.3 mL of concentrated HCl (12 M) to 1.00 liter 
pH = 3 solution: dilute 10. mL of pH = 1 solution (above) to 1.00 liter 
pH = 5 solution: dilute 10. mL of pH = 3 solution (above) to 1.00 liter 
pH = 7 solution: distilled water 

pH = 13 solution: dissolve 4.00 g of NaOH in sufficient water to produce 1.00 liter of solution 
pH = 11 solution: dilute 10. mL of pH = 13 solution (above) to 1.00 liter 
pH = 9 solution: dilute 10. mL of pH = 11 solution (above) to 1.00 liter 

methyl orange indicator: dissolve 0.1 g of methyl orange powder in 100 mL of water and filter 
bromthymol blue indicator: dissolve .01 g of bromthymol blue in 100 mL of 50% water and 50% 
ethanol solution and filter 
• phenolphthalein solution: dissolve 1.0 g of phenolphthalein powder in 100 mL of ethanol 

To prepare the unknown solutions for Part IV, select three of the known pH solutions used in the lab and 
label them as unknown. Be sure to keep a record of which pH values were selected as unknowns. 

Answers to Pre-Lab Questions 

1. How is universal indicator made? 
Several indicators are mixed. 

2. What distinguishes weak organic acids that are useful as acid-base indicators from weak organic acids 
that will not function as acid-base indicators? 

The undissociated molecule of the acid and the anion of the dissociated acid must be different colors. 

pH Measurements Using Indicators 

Background Information 

The nature of acids and bases have been known to man for quite sometime. Chemically speaking, acids 
are interesting compounds because a large number of common household substances are acids or acidic 
solutions. For example, vinegar contains ethanoic acid, also called acetic acid, //C2//3O2) and citrus fruit 
contain citric acid. Acids cause foods to have a sour taste and turn litmus red. (Note: You should never 
taste substances in the laboratory.) Also, many common household substances are bases. Milk of magnesia 
contains the base magnesium hydroxide, Mg{OH)2 and household ammonia is a common cleaning agent. 
Bases have a slick feel to the fingers and turn litmus blue. (Note: You should never feel chemicals in the 
laboratory.) 

Indicator dyes, of which litmus is one, turn various colors according to the strength of the acid or base 
applied to it. 

Pure water, which is neutral in terms of acid-base, exists mostly as H2O molecules but does, to a very 
slight extent dissociate into hydrogen and hydroxide ions. 

537 www.cki2.0rg 



The extent of this dissociation is 1.0x10 7 moles /liter (at 25°C). Therefore, in all neutral water (and neutral 
water solutions), the concentration of hydrogen ions is 1.0xl0~ 7 M and the concentration of hydroxide ions 
is 1.0 x 10~ 7 M. The dissociation constant for this process is K w = [H + ][OH~] = (1.0 x 10~ 7 )(1.0 x 10~ 7 = 
1.0 xlO" 14 ). 

In 1909, a Danish chemist (Soren Sorenson), developed a mathematical system for referring to the degree 
of acidity of a solution. He used the term pH for "power of hydrogen" and established the equation, 
P H = -log[H+}. 

In a neutral solution, the hydrogen ion concentration is 1.0 X 10~ 7 M and therefore, the pH is 7. If the 
concentration of hydrogen ions is 1.0 X 10 -5 M, then the pH is 5. A solution is neutral when the pH equals 
7, it is acid if the pH is less than 7, and it is basic if the pH is more than 7. In commonly used solutions, 
pH values usually range from 1 to 14. 

Living matter (protoplasm) contains a mixture of variously dissociated acids, bases, and salts and usually 
has a pH very near neutral. The pH of human blood is generally 7.3 and humans cannot survive if the 
blood becomes more basic than pH 7.8 or more acidic than pH 7.0. Life of any kind exists only between 
pH 3 and pH 8.5. Buffer solutions regulate the pH of the body by neutralizing excess acid or base. The 
chief buffers of the body are proteins, carbonates, phophates, and hemoglobin. The kidneys play a role by 
eliminating excess electrolytes. 

Table 18.2: Some Typical 



pH 



Substance 



Acidity /Basicity 




1 
2 

3 

4 

5 

6 

7 

8 

9 

10 

11 

12 

13 

14 



Sulfuric Acid (Battery Acid) 
0.10 M Hydrochloric Acid 
Stomach Acid 

Vinegar 

Tomato Juice 

Black Coffee and Vitamin C 

Cow's Milk 

Distilled Water 

Sea Water 

Baking Soda, NaHCOj, 

Detergents 

Household Cleaning Ammonia 

0.10 M NaOH 

1.0 M NaOH (Lye) 



Very Highly Acidic 
Highly Acidic 
Acidic 



Acidic (jj^' 



as strong as pH 1] 



Weakly Acidic 

Very Weakly Acidic 

Neutral 

Very Weakly Basic 

Weakly Acidic 

Basic 

Strongly Basic 
Very Strongly Basic 



Some acids dissociate completely into ions when dissolved in water. Such acids are called strong acids 
(HCl, HI, HBr, HNO%, H2SO4, HCIO4). Some bases dissociate completely when dissolved in water. Such 
bases are called strong bases (NaOH, LiOH, KOH, RbOH). There are other acids and bases that dissociate 
only slightly (although completely soluble) when dissolved in water. Such acids and bases are called weak 
acids or weak bases and some examples are HF , HC2H3O2, NH4OH. 

An important method of determining pH values in the lab involves the use of substances called "acid-base 
indicators". These are certain organic substances (almost always weak organic acids) that have the property 
of changing color in solutions of varying hydrogen ion concentration. In order for a weak organic acid to 
be useful as an acid-base indicator, it is necessary that the undissociated molecule and the indicator anion 
be different colors. For example, phenolphthalein is a colorless substance in any aqueous solution in which 
the hydrogen ion concentration is greater than 1 X 10 -9 M (pH < 9) but changes to a red or pink color 



www.ckl2.org 



538 



when the hydrogen ion concentration is less than 1 x 10 -9 M (pH > 9). Such substances can be used for 
determining the approximate pH of solutions. Electrical measurements can determine the pH even more 
precisely. This lab will use three acid base indicators and what is called a "universal indicator". 

Table 18.3: Some Indicator Color Changes 

Indicator pH Color Change Range Color Change 

Methyl Orange 3.1 - 3.4 Red to Yellow 

Bromthymol Blue 6.0 - 7.6 Yellow to Blue 

Phenolphthalein 8.3 - 10.0 Colorless to Red 

The universal indicator (one type is called Bogen's Universal Indicator), is made by mixing a number of 
indicators that all change color at different pH's. As you slowly change the pH of the indicator from 1 
to 14, it goes through a series of subtle color changes. The indicator is provided with a photographic 
chart that shows the color of the indicator at every different pH and the pH is identified by matching the 
indicator color to the chart. 

Pre-Lab Questions 

1. How is universal indicator made? 

2. What distinguishes weak organic acids that are useful as acid-base indicators from weak organic acids 
that will not function as acid-base indicators? 

Purpose 

The purpose of this lab is to have the student experience the color changes involved with acid-base indicators 
and to identify the approximate pH of an unknown solution using acid-base indicators. 

Apparatus and Materials 

Well Plates, at least 12 wells (1 per lab group) 

drop controlled bottles of pH = 1 

drop controlled bottles of pH = 3 

drop controlled bottles of pH = 5 

drop controlled bottles of pH = 7 

drop controlled bottles of pH = 9 

drop controlled bottles of pH = 11 

drop controlled bottles of pH = 13 

drop controlled bottles of methyl orange 

drop controlled bottles of bromthymol blue 

drop controlled bottles of phenolphthalein 

drop controlled bottles of universal indicator 

drop controlled bottle of unknown #1 

drop controlled bottle of unknown #2 

drop controlled bottle of unknown #3 

Safety Issues 

All solutions are irritating to skin, eyes, and mucous membranes. Handle solutions with care, avoid getting 
the material on you, and wash your hands carefully before leaving the lab. 

Procedure for Part I: Determining the effect of pH on indicator dyes. 

539 www.ckl2.org 



1. Place the Chemplate on a sheet of white paper. 

2. Place one drop of methyl orange into cavities #1 and #2. 

3. Place one drop of bromthymol blue in cavities #5 and #6. 

4. Place one drop of phenolphthalein in cavities #9 and #10. 

5. Carefully add one drop of pH 1 to cavities #1, #5, and #9. 

6. Carefully add one drop of pH 13 to cavities #2, #6, and #10. 

Data for Part I 



1. What color is the original methyl orange solution? 

2. What color is methyl orange in a strong acid? 



3. What color is methyl orange in a strong base? 

4. What color is the original bromthymol blue solution? 

5. What color is bromthymol blue in a strong acid? 

6. What color is bromthymol blue in a strong base? 



7. What color is the original phenolphthalein solution? 

8. What color is phenolphthalein in a strong acid? 

9. What color is phenolphthalein in a strong base? 

Rinse the Chemplate in tap water and dry with a paper towel. 

Procedure for Part II: Determining the pH color change range of indicator dyes. 

1. Place one drop of methyl orange in each cavity numbered 1-7. 

2. Carefully add one drop of pH 1 to cavity #1, pH 3 to cavity #2, pH 5 to cavity #3, pH 7 to #4, 
pH 9 to #5, pH 11 to #6, and pH 13 to #7. 

3. Repeat the rinsing, drying, and steps 1 and 2 except using bromthymol blue and then repeat the 
entire process again using phenolophthalein. 

Data for Part II 

1. Describe the color changes and the pH 's around the color change pH for methyl orange. 

2. Describe the color changes and the pH 's around the color change pH for bromthymol blue. 

3. Describe the color changes and the pWs around the color change pH for phenolphthalein. 

Rinse the Chemplate in tap water and dry with a paper towel. 

Procedure for Part III: Determining a color standard for universal indicator. 

1. Place one drop of universal indicator in each cavity numbered 1-7. 

2. Carefully add one drop of pH 1 to cavity #1, pH 3 to cavity #2, pH 5 to cavity #3, pH 7 to #4, 
pH 9 to #5, pH 11 to #6, and pH 13 to #7. 

Keep these solutions for Part IV. 
Data for Part III 

1. Describe the color of the universal indicator at each pH used. 

Cavity #1 (pH — 1), color = 

Cavity #2 (pH = 3), color = 

www.ckl2.org 540 



Cavity #3 (pH = 5), color = 
Cavity #4 (pH = 7), color = 
Cavity #5 (pH = 9), color = 
Cavity #6 (pH =11), color 
Cavity #7 (pH = 13), color 



Procedure for Part IV: Determining the pH of some unknown solutions. 

1. Place one drop of universal indicator in cavities #10, #11, and #12. 

2. Place one drop of unknown #1 in cavity #10. 

3. Place one drop of unknown #2 in cavity #11. 

4. Place one drop of unknown #2 in cavity #12. 

5. Compare the color in each cavity with the colors in cavities #1-7 that you made in activity 3. 

Data for Part IV 

Color of unknown #1 in universal indicator 

Color of unknown #2 in universal indicator 

Color of unknown #3 in universal indicator 

Post-Lab Questions 

1. What is the pH of unknown #1? 

2. What is the pH of unknown #2? 

3. What is the pH of unknown #3? 

18.3 Titrations 

Lesson Objectives 

Define titrations and identify the different parts of the titration process. 

Explain the difference between the endpoint and the equivalence point. 

Describe the three types of titration curves. 

Identify points on the titration curves for the three types of titrations. 

Define a standard solution. 

Calculate the accurate concentration of an acid or base using a standard. 

Calculate unknown concentrations or volumes of acids or bases at equivalence. 

Introduction 

For acid-base neutralization reactions, the typical laboratory procedure for determining the stoichiometric 
amounts of acid and/or base in the reaction is to complete a titration. There are three main types of 
titration experiments and each of these has some similarities but also have some differences. A titration 
experiment can involve equipment from the simple (such as using an eyedropper) to the complex (such as 
using a burette, a pH meter, a magnetic stirrer). As we go through this lesson, we will use some of the 
prior knowledge we have obtained about acids and bases, about chemical reactions, molarity calculations, 
and about indicators to apply them to the concept of titrations. 

541 www.ckl2.org 



Titration 

One of the properties of acids and bases is that they neutralize each other. In the laboratory setting, 
an experimental procedure where an acid is neutralized by a base (or vice versa) is known as a titration. 
A titration, by definition, is the addition of a known concentration of base (or acid) (also called the 
titrant) to a solution of acid (or base) of unknown concentration. Since both volumes of the acid and base 
are known, the concentration of the unknown solution is then mathematically determined. So what does 
one do in a titration? When doing a titration, you need to have a few pieces of equipment. A burette 
(Figure ??) is used to accurately dispense the volume of the solution of known concentration (either the 
base or the acid). An Erlenmeyer flask is used to hold a known volume of the unknown concentration 
of the other solution (either the acid or the base). Also an indicator is used to determine the endpoint 
of the titration. A few drops of the indicator are added to the flask before you begin the titration. The 
endpoint is the point where the indicator changes color, which tells us that the acid is neutralized by the 
base. The equivalence point is the point where the number of moles of acid exactly equals the number 
of moles of base. The equivalence point is a calculated point in the neutralization of the acid and the base. 
Some laboratories have pH meters (Figures 18.5, 18.7) that measures this point more accurately than 
the indicator although an indicator is much more visual! Figure 18.5 shows a simplified version of a pH 
meter with the probe from the meter immersed in the liquid. Figure 18.6 shows a typical electronic pH 
meter with the attached probes. The main purpose of a pH meter is to measure the changes in pH as the 
titration goes from start to finish. 

Burette. (Source: http://commons.wikimedia.Org/wiki/File:Burette.png. Public Domain) 




Shown above is a typical titration setup. The burette is upright ready to drip the solution into the flask 
holding the solution of unknown concentration and the few drops of indicator. When the indicator changes 
color, the number of moles of acid equals the number of moles of base and the acid (or base) has been 
neutralized. 

There are three types of titrations that are normally performed in the laboratory in order to determine 
the unknown concentration of the acid or base. These three types are: 

1. Strong acid vs. Strong base 
www.ckl2.org 542 










• 


8.03 






• • 





Figure 18.5: A simple meter with its probe immersed in a mildly alkaline solution . The two knobs on the 
meter are used to calibrate the instrument. 




Figure 18.6: An electronic meter 



543 



www.ckl2.org 



■_ 
: 



: 
7 




Figure 18.7: Typical Titration Setup. 



www.ckl2.org 



544 



2. Strong acid vs. Weak base 

3. Weak Acid vs. Strong base 

In these titrations, a pH meter may be used to measure the changes in the pH as the titration goes to 
completion. If so, a titration curve can be constructed for each curve. A titration curve is a graph of the 
pH versus the volume of titrant added. Let's take a look at how each of these types of titrations differs in 
terms of their pH curves and their pH at the equivalence point. 

(1) Strong Acid vs. Strong Base 

For a strong acid vs. a strong base titration, let's assume the strong base is the titrant. Therefore, in 
the Erlenmeyer flask is the strong acid and a few drops of your indicator. Think about the pH of the 
solution in the flask. Do you think it will be high (around pH = 11.0), mildly basic (around pH = 8.0), 
mildly acidic (around pH = 6.0), or low (around pH = 1.0)? Probably the pH will be around 1.0 since the 
solution is a strong acid. As the base is added, the acid is slowly neutralized. At first the change in pH is 
minimal. This resistance is due to the fact that the flask has a much greater number of H^O + ions than 
the OH~ ions available from the added titrant. 

As more and more base is added, more OH ions are added and thus more H%0 + ions get neutralized. 
Let's stop here and look at the reaction. Equation 1 shows the total ionic equation of a reaction between 
a strong acid and a strong base. 



H+ (aq) + CI (aq) + Na+ (aq) + OH ( ag ) -» NcT \ aq ) + CI ( aq ) + H 2 0( L ) 



(Equation 1) 



Equation 2 shows the net ionic equation for the reaction between the strong acid and the strong base. 
H + {aq ) + OH- {aq) -> H 2 {L) (Equation 2) 

So you see that as we add more OH~ ions, more H^O + (or H + ) ions are being neutralized. Since these two 
ions react to form water, a neutral solution will be formed. For a strong acid and a strong base, this means 
the pH = 7.0 at the point of neutralization. If we continue to add the titrant (containing O// - ions) after 
all of the H^O + ions have been neutralized, the pH will continue to rise as more base is added and there 
are excess OH~ ions. 

Now we know what happens in a strong acid/strong base titration, what does the titration curve look like? 
Look at Figure 18.8. The main points of the titration curve described above are shown in the titration 
curve below. 




D Equivalence point 



Volume of titrant added 

Figure 18.8: Titration Curve for a Strong Acid vs. Strong Base. 

The points A through D sum up the description of the events that take place during the titration. Point 
A is the start of the titration. Point B is the midpoint, the point where half of the H + ions have been 
neutralized. Point D is the equivalence point. 



545 



www.ckl2.org 



(2) Strong Acid vs. Weak Base 

What would happen if we were to titrate a strong acid with a weak base or vice versa? Look at Figure 
18.9. Can you determine what is happening in the titration just by looking at the graph? 



pH 




Equivalence point 



Volume of titrant added 



Figure 18.9: Titration Curve for a Strong Acid vs. Weak Base. 

When you look at Point A, the pH is high but not too high {pH approx 11). This means that there is a 
weak base in the flask. As the acid (the titrant) is added, the pH decreases as the H 3 + ions being added 
and begin to neutralize the OH~ ions. Point C represents the point where the volume of acid necessary to 
neutralize the base is measured. Therefore, if we draw a line down to the x-axis, we can find the volume of 
titrant necessary for the neutralization reaction (same as before). Point D is the equivalence point. Notice 
that for a weak base and a strong acid titration, the pH at equivalence point is more acidic. Equation 3 is 
the reaction between NH 



3(aq)i a weak base, and HClr aq \, a strong acid. 

NH 3(aq) + HCl(aq) "» ^ H 4 C l(aq) + #20 (L) 

The ionic equation is Equation 4. 

NH 3(aq) + h+ cj) + cr (uq) -> NH 4 + {aq) + cr {aq) + H 2 {L) 



(Equation 3) 



(Equation 4) 



(3) Weak Acid vs. Strong Base 

The third type of titration is that of a weak acid with a strong base. When we follow through with the 
same procedure as was done with the previous two titrations, we can determine a great deal of information 
simply by looking at the pH curve. For example, let's consider the titration of a solution of acetic acid, 
HC2H3O2, with a solution of potassium hydroxide, KOH. We can write the chemical reaction for this 
acid-base neutralization (see Equation 5) and begin to draw a rough sketch of a titration curve. 



H+ (aq) + C 2 H 3 2( + K + {aq) + OH {aq) -> K + {aq) + C 2 H 3 2( + H 2 {L) 



(Equation 5) 



Acetic acid is a weak acid and we know that the pH at the start of the titration will be around 2.8. 
Potassium hydroxide is a strong base so the curve will end high on the graph, around a pH = 12. 




p. Equivalence point 



Volume of titrant added 



The points on the curve represent the same points as with the other two titration curves. Look, however, at 
the equivalence point. Notice how the pH for the equivalence point of the weak acid- strong base titration 
is above 7.0. Look at Equation 5. 



www.ckl2.org 



546 



Sample question: Draw a rough sketch of the titration curve between nitric acid and ethylamine, CH3NH2. 
Assume the acid is in the burette. What is the estimated pH at the equivalence point? 

Solution: 




Volume of titrant added 



The pH at the equivalence point is approximately 4.6 from this graph. 

Standard Solutions 

As was previously stated, when we do a titration the titrant is the solution of known concentration. For 
accuracy reasons, this titrant is normally titrated to find the exact concentration before beginning the 
acid-base titration. The purpose of this initial titration is to determine, with as much accuracy as possible, 
the exact concentration of the solution in the burette. To determine the exact concentration of the titrant, 
we use a standard solution. A standard solution is a solution whose concentration is known exactly. 
Standard solutions have this property because these chemicals are normally found in pure, stable forms. 
Examples of chemicals used to prepare standard solutions are the acidic potassium hydrogen phthalate, 
KHC 7 H^O^ (sometimes referred to as KHP), and the basic sodium carbonate, NCI2CO3. 

When using a standard solution, the standard is first prepared by dissolving the solid in a known volume 
of water, add a few drops of indicator, and titrate with the solution that you want to standardize. 

Sample question: What is the concentration of sodium hydroxide when 32.34 mL is required to neutralize 
a solution prepared by dissolving 1.12 g of KHC 7 H/^0^ in 25.00 mL of H^Ofj^l 

Solution: 

Step 1: Find the mols of KHC 7 H^O^t s y 

mass 1.12 g , 

mols KHC 7 H 4 4 = — = f- — - = 5.83 x 10~ 3 mol 

molar mass 192.2 g/mol 

Step 2: Use mol ratio from the reaction to find the moles of NaOH. 

KHC 7 H±0 A{aq) + NaOH {aq) -> KNaC 7 H 4 4{aq) + H 2 {L) 

Since the reaction is 1:1, or 1 mole of KHP reacts with every mole of NaOH, the number of moles of 
KHP = number of moles of NaOH. 

mol NaOH = 5.83 X 10~ 3 mol 

Step 3: Determine the concentration of NaOH. 

. 5.83 x 10~ 3 mol 

NaOH] = = 0.180 M 

L J 0.03234 L 

Therefore, the exact concentration of the sodium hydroxide solution used in the titration is 0.180 mol/L. 

547 www.ckl2.org 



Choosing an Appropriate Indicator 

To choose an appropriate indicator for a titration, a titration curve is useful. Knowing the pH at equivalence 
for the different types of titrations (see Table 18.4) is also needed. 

Table 18.4: pH at Equivalence for Titrations 



Type of Titration 



pH at Equivalence 



Strong Acid - Strong Base 
Strong Acid - Weak Base 
Weak Acid - Strong Base 



pH = 7.0 
pH < 7.0 
pH > 7.0 



Choosing an indicator close to the equivalence point is essential to see the point where all of the H + ions 
and OH" ions have been neutralized. The color change should occur on or around the equivalence point. 
So, for example, with a strong acid, strong base titration, the pH at equivalence is 7.0. Indicators such 
as broniothymol blue [pH range = 6.0 - 7.6) and phenol red {pH range = 6.6 - 8.0) are common. If you 
notice the midpoint color (green) for broniothymol blue would appear at a pH = 6.8 which is close to 7.0. 
For phenol red, the midpoint color (orange) would appear at pH = 7.3, again close to 7.0. 

The same process is used for other titration types. For a strong acid-weak base titration where the pH at 
equivalence is less than 7, the indicators normally chosen for these titrations are methyl red {pH range = 
4.8 — 6.0) and chlorophenol red {pH range = 4.8 - 6.2). For a weak acid-strong base titration where the pH 
at equivalence is greater than 7, the indicators normally chosen for these titrations are phenolphthalein 
{pH range = 8.2 - 10) and thymol blue {pH range = 8.0 - 9.6). As with strong acid-strong base titration, 
the visual observation of the midpoint color should indicate close proximity to the equivalence point. 

Sample question: Look at the graph below and determine the appropriate indicator. 

Graph 1: Titration of a Weak Acid with a Strong Base (Source: Therese Forsythe. CC-BY-SA) 




Solution: 

We first look at the graph and mark the vertical stretch of the titration curve (red lines) in order to find 
the half-way mark on this vertical stretch (green line). Looking at the graph, when we follow this half-way 
mark over to the y-axis, we can see that the equivalence point occurs at approximately pH = 8.2. The 
indicator appropriate to use would be phenolphthalein {pH range = 8.2 - 10) and as soon as the pink color 
forms we are at the equivalence point. 

Graph 2: Titration of a Weak Acid with a Strong Base (Source: Therese Forsythe. CC-BY-SA) 



www.ckl2.org 



548 




Reaction Continues to the End Point 

There is an interesting observation about the endpoint that has yet to be mentioned. The endpoint was 
defined earlier as the point where the indicator changes color. In an acid-base neutralization reaction, this 
point may not be the point where all of the H + ions have been neutralized by OH" ions, or vice versa. The 
experimenter continues titration until the indicator changes color, that is, the endpoint has been reached. 
The equivalence point is the point where the moles of hydrogen ion and the moles of hydroxide ion are 
equal. It requires knowledge by the experimenter to select an indicator that will make the endpoint as 
close as possible to the equivalent point. 

Titration Calculations 

For the calculations involved here, we will restrict our acid and base examples where the stoichiometric 
ratio of H + and OH~ is 1 : 1. To determine the concentration or volume required to neutralize and acid or 
a base, in other words, to reach the equivalence point, we will use a formula similar to the dilution formula 
used in your prior learning. The formula has the structure: 

M a x V a = M b x V h 

— where — 

M a = molarity of the acid 

V a = volume of the acid 
Mf, = molarity of the base 

Vb = volume of the base 

Sample question: When 10.0 mL of a 0.125 mol/L solution of hydrochloric acid, HCl, is titrated with a 
0.100 mol/L solution of potassium hydroxide, KOH, what is the volume of the hydroxide solution required 
to neutralize the acid? What type of titration is this? 

Solution: 

Step 1: Write the balanced ionic chemical equation. 



H (aq) + CI ( m/ ) + K ( ag ) + OH ( ag ) -> K ( a? ) + CI ( a? ) + H 2 0( L -) 



Step 2: Use the formula and fill in all of the given information. 

549 



www.ckl2.org 



M a x V a = M h x V h 

M a = 0.125 mol/L 

V a = 10.0 mL 

M h = 0.100 mol/L 

V* = ? 

M a xV a = M b x V fo 

M fl xV fl (0.125 mol/L) (10.0 mL) 

K/, = = ; = 12.5 mL 

M b 0.100 mol/L 

Therefore, for this weak acid-strong base titration, the volume of base required for the titration is 12.5 mL. 

Lesson Summary 

• A titration is the addition of a known concentration of base (or acid) to a solution of acid (or 
base) of unknown concentration. The titrant in the titration is the solution of known concentration. 
This solution is normally in the burette. A burette is a piece of equipment used in a titration to 
accurately dispense the volume of the solution of known concentration (either the base or the acid) . 
The endpoint is the point in the titration where the indicator changes color. The equivalence point 
is the point in the titration where the number of moles of acid equals the number of moles of base; 
pH meters are used in many laboratories for acid-base titrations. The main purpose of a pH meter 
is to measure the changes in pH as the titration goes from start to finish. 

• The three types of titrations usually performed in the laboratory are: strong acid vs. strong base, 
strong acid vs. weak base, and weak acid vs. strong base. A titration curve is a graph of the pH 
versus the volume of titrant added. For a strong acid vs. strong base titration, the pH at equivalence 
is 7.0. For a strong acid vs. weak base titration, the pH at equivalence is less than 7.0. For a 
weak acid vs. strong base titration, the pH at equivalence is greater than 7.0. A standard solution 
is a solution whose concentration is known exactly and is used to find the exact concentration of 
the titrant. For titrations where the stoichiometric ratio of mol H + : mol OH~ is 1 : 1, the formula 
M a x V a = Mb x Vb can be used to calculate concentrations or volumes for the unknown acid or base. 

Review Questions 

1. Why do you think there would be more experimental error with using an indicator over using a pH 
meter in a titration? 

2. Why would there not be a weak acid- weak base titration? 

3. Which of the following definitions best suits that of an endpoint? 

(a) The stoichiometric point where the number of moles of acid equals the number of moles of base. 

(b) The visual stoichiometric point where the number of moles of acid equals the number of moles 
of base. 

(c) The midpoint of the vertical stretch on the titration curve. 

(d) None of the above 

4. In the following titration curve, what pair of aqueous solutions would best represent what is shown 
to be happening in the curve? 

(a) HCOOH (aq) + NH 3{aq) 
www.ckl2.org 550 



(b) HCOOH (aq) + NaOH (aq) 

(c) H 2 SO Haq) +Ba(OH) 2{aq) 

I A\ Ufln . , . 1 Mil- , . 



(d) HClO A{aq) + Wtt; 



3 (a?) 




1 
80 



i i r 

20 40 

Volume(mL) 

5. What would be the best indicator to choose for the pH curve shown in question 3? 

(a) Methyl red 

(b) Litmus 

(c) Phenolphthalein 

(d) Phenol red 

6. What is the best indicator to use in the titration of benzoic acid with barium hydroxide? 

(a) Methyl violet 

(b) Bromothymol blue 

(c) Phenolphthalein 

(d) Methyl blue 

(e) Indigo carmine 

Table 18.5: 



Indicator 



pH Range 



Methyl violet 
Bromothymol blue 
Phenolphthalein 
Methyl blue 
Indigo carmine 



0.0-1.6 
3.0-4.7 
8.2 - 10.0 
10.6-13.4 
11.4-13.0 



7. If 22.50 mL of a sodium hydroxide is necessary to neutralize 18.50 niL of a 0.1430 mol/L HNO3 
solution, what is the concentration of NaOHl 

(a) 0.1176 mol/L 

(b) 0.1430 mol/L 

(c) 0.1740 mol/L 

(d) 2.64 mol/L 

8. Plot the following titration data on a titration curve of pH vs. volume of base added. When complete, 
find the pH at equivalence and choose an appropriate indicator for the titration. What volume of 



551 



www.cki2.0rg 



base is necessary to neutralize all of the acid? 



Table 18.6: 



Volume of base added (mL) pH 



0.00 1.0 

2.00 1.2 

4.00 1.4 

6.00 1.6 

8.00 1.9 

9.00 2.3 

9.50 2.6 

9.90 3.3 

9.99 4.3 

10.00 7.0 

10.01 9.7 
10.50 10.7 
12.00 11.4 
14.00 12.1 
16.00 12.1 



9. Calculate the concentration of hypochlorous acid if 25.00 mL of HCIO is used in a titration with 
32.34 mL of a 0.1320 mol/L solution of sodium hydroxide. 

Further Reading / Supplemental Links 

http : //en . wikipedia . org 

Vocabulary 

titration The lab process in which a known concentration of base (or acid) is added to a solution of acid 
(or base) of unknown concentration. 

titrant The solution in the titration of known concentration. 

burette A piece of equipment used in titrations to accurately dispense the volume of the solution of 
known concentration (either a base or an acid). 

Erlenmeyer flask A piece of equipment used in titrations (and other experiments) to hold a known 
volume of the unknown concentration of the other solution (either the acid or the base). 

endpoint The point in the titration where the indicator changes color. 

equivalence point The point in the titration where the number of moles of acid equals the number of 
moles of base. 

pH meter A device used to measure the changes in pH as the titration goes from start to finish. 

www.ckl2.org 552 



titration curve A graph of the pH versus the volume of titrant added. 

standard solution A solution whose concentration is known exactly and is used to find the exact con- 
centration of the titrant. 

Labs and Demonstrations for Titrations 
Teacher's Pages for Acid-Base Titration 



A 



Investigation and Experimentation Objectives 



In this activity, the student will use burets, pipets, and acid-base indicators to collect data. The properly 
recorded data with then be used along with several mathematical formulas to determine unknown molecular 
masses from titration data. 

Lab Notes 

Students will need two days to do both parts of the lab. 

Preparation of Solutions, KHP, and unknown acids 

6 M NaOH solution: Boil 600 mL of distilled water to drive off any dissolved COi- (The CO2 produces 
carbonic acid, which drives down the concentration of NaOH.) Add 120. g of NaOH to a 500 mL volumetric 
flask, and add some of the freshly boiled distilled H20 to the flask. Swirl to dissolve. Cool the resultant 
solution in a cold water or ice water bath, let the solution and flask return to room temperature, and dilute 
the resulting solution to 500 mL. Store this solution in a tightly capped bottle, preferably Nalgene or other 
base-resistant bottle. This will provide enough solution for (75) two-student teams with a 20% excess for 
spills and endpoint over-runs. 

Phenolphthalein indicator solution: Dissolve 0.1 g of phenolphthalein in 50 mL of 95% ethanol, and dilute 
to 100 mL by adding distilled water. This will provide enough solution for (75) two student teams. 

KHP - Potassium Hydrogen Phthalate - Dry 200 g of KHP in a laboratory oven for at least one hour prior 
to titration. Store the KHP in a dessicator. KHP is slightly hygroscopic. This is sufficient KHP for (75) 
two - student teams. 

Unknown Acids 

The following acids are suggestions for use. Their number of ionizable hydrogens vary, are stable chemically, 
and many schools have them on hand. Twenty grams of each acid are required for (75) two - student teams: 

Table 18.7: The Number of Ionizable Hydrogens and Molar Mass of Various Acids 

Acid Number of Ionizable Hydro- Molar Mass (g/mol) 

gens 

Lactic 1 90.1 

Malonic 2 104.1 

Maleic 2 116.1 

Succinic 2 118.1 

Benzoic 1 122.1 

Salicylic 1 138.1 

Tartaric 2 150.1 

553 www.cki2.0rg 



NOTE: The molecular weights listed are for the anhydrous acids. If hydrates are used, use the molecular 
weights as shown on the reagent bottle. The number of ionizable hydrogens does not change. 

Answers to Pre-Lab Questions 

1. Pre-rinsing the buret will remove any water or residual NaOH solution within the buret. If there were 
water present, it would dilute the NaOH added. If there were residual NaOH, the concentration would 
increase upon addition of solution due to the crystalline or concentrated solution of NaOH present. 

2. The KHP and the NaOH react with each other in a 1:1 stoichiometric ratio. Thus, the number of moles 
of NaOH required to react with the KHP will be equal to the number of moles of KHP originally present. 

Moles of NaOH added = M x L = (0.1 M)(0.040 L) = 0.004 moles NaOH added: thus 

Moles KHP needed = 0.004 moles 
Grams of KHP needed = (mole s)(g/ mole) = (0.004 mot) (204.23 g/mol) = 0.8 grams KHP 

3. Molarity of KHP - moUs KHP 



liters of solution 



Moles of KHP = JZZfKHP = 2oI?fc = °-°° 372 m ° l KHP 

Molarity = 0003 7 g 5 ^ / / //p = 0.0743 M KHP 

4. Again, the amount of water needed to dissolve the KHP is not needed to solve this problem. All you 
need is the weight of the KHP and it's molecular weight : 

Moles of KHP = ' i— = 0.00255 mols KHP 

J 204.2 g/mol 

Since this will be equal to the number of moles of NaOH reacted, the number of liters of titrant you need 
to add can be calculated by re-arranging the molarity formula: 

moles 

M = 

L 

moles 0.00255 mol 

so L = = — = 0.0213 L = 21.3 mL 

M 0.102 mol/L 

Acid-Base Titration Lab 

Background Information 

Standardization of the sodium hydroxide solution through titration is necessary because it is not possible 
to directly prepare a known molarity solution of sodium hydroxide with high accuracy. Solid sodium 
hydroxide readily absorbs moisture and carbon dioxide from the atmosphere and thus it is difficult to 
obtain a precise amount of the pure substance. A sodium hydroxide solution will be made close to 0.1 M 
and then the actual molarity of the solution will be determined by titration of a primary standard. A 
primary standard is a substance of very high purity that is also stable in air. Because the substance 
remains pure, it is possible to mass a sample of the substance with a high degree of accuracy. The primary 
standard used in this experimental procedure is potassium hydrogen phthalate, KHCgH^O^ (molar mass 
204.2 g j mole). 

This primary standard, KHP, is used to standardize the secondary standard, sodium hydroxide. The 
standardized sodium hydroxide solution can then be used to determine the molar mass of an unknown acid 
through titration. 

In both steps a titration is performed in which a buret is used to dispense measured increments of the sodium 
hydroxide solution into a second solution containing a known mass of KHP (NaOH standardization). For 

www.ckl2.org 554 



the second reaction, a mass of acid whose molecular weight is unknown is then titrated with the solution 
of NaOH whose concentration was determined by the standardization with KHP. The stoichiometry of the 
reaction depends on the number of ionizable hydrogens within the acid. KHP is a weak monoprotic acid 
that will react with sodium hydroxide in a 1:1 mole ratio: 

KHC 8 H A 4 + NaOH -> KNaC 8 H 4 A + H 2 

The unknown acid may be monoprotic, diprotic, or triprotic dependent on the number of acidic hydrogens 
present in the molecule. A monoprotic acid, HA, has one acidic hydrogen, a diprotic acid, H 2 A, two acidic 
hydrogens, and a triprotic acid, H3A, three acidic hydrogens. The stoichiometries of the reactions are 
shown below. You will be told whether your unknown acid is monoprotic, diprotic, or triprotic. 

Monoprotic HA + NaOH -> NaA + H 2 
Diprotic H 2 A + 2NaOH -^ Na 2 A + 2H 2 
Triprotic H 3 A + ZNaOH -^ Na 3 A + 3H 2 

The indicator phenolphthalein is used as a signal of the equivalence point. Phenolphthalein is a weak 
organic acid that will change from colorless to pink near the equivalence point of the titrations. The actual 
point at which the indicator changes color is the end point. The endpoint and the equivalence point are 
not the same. The difference between the two is the titration error. Obviously, for a titration to be of 
value, care must be taken to select an indicator for which the difference between the equivalence point and 
the endpoint is small. With this particular titration, it is very small. 

Pre-Lab Questions 

1. Why is it necessary to rinse out the buret with the NaOH solution? 

2. Calculate the approximate weight of KHP required so that about 40 mL of 0.1 M sodium hydroxide 
is used in a titration. (MW of KHP = 204.23 g/mol) 

3. Calculate the molarity of a KHP solution when 0.759 g of KHP is dissolved in 50.0 mL of water. 

4. 0.521 g of KHP is dissolved in 40 mL of water, and titrated with a 0.102 M NaOH solution. Calculate 
the number of mL of the NaOH solution added. 

Purpose 

The purpose of this experiment is to determine the concentration of a titrating solution, NaOH, using a 
stable compound, KHP. Once the concentration of the NaOH solution is known, it can then be used to 
determine the molecular weight of an acid whose formula is unknown. 

Apparatus and Materials 

• 50 mL buret 

• buret stand 

• buret clamp 

• 125 mL Erlenmeyer flask 

• phenolphthalein 

• NaOH solution 
. KHP, solid 

• unknown acid, solid 

• 10 mL graduate 

• 400 mL beaker 

• 100 mL graduate 

555 www.cki2.0rg 



Safety Issues 

NaOH is a caustic solution and will cause severe burns, especially to eye tissue. Wear goggles and aprons 
at all times. The solid acids cause considerable irritation if exposed to skin or mucous membranes. Avoid 
exposure. 

Procedure for Part I 

Part 1. Standardization of Sodium Hydroxide Solution 

Obtain the primary acid standard from your instructor. Record the name of the acid, its molecular 
formula, and number of acidic hydrogens per molecule. Prepare about 300 mL of approximately 0.1 M 
sodium hydroxide by diluting 6 M NaOH with distilled water. (Calculate, ahead of time, how much water 
and how much 6 M NaOH will be needed.) 

WARNING: Concentrated sodium hydroxide is corrosive and causes severe burns. Handle with care. 
Dilute and wash up spills with plenty of water. Wash affected skin with water until it no longer feels 
slippery, but feels "squeaky" clean. 

Store the solution in your plastic bottle, label it "0.1 M NaOH", and keep it tightly capped. You will 
determine the exact molarity of this NaOH solution by standardization. 

Calculate the mass of the primary acid standard that would react with about 20 mL of 0.1 M NaOH. 
Weigh approximately this amount into a clean 125 - mL Erlenmeyer flask by taring the balance with the 
flask on the pan, and then adding the acid to the flask. Record the mass of the primary acid standard 
to the highest precision allowed by the electronic balance. Add 30 to 40 mL of your purified water to the 
flask, and swirl to dissolve the primary acid standard. Add three or four drops of phenolphthalein indicator 
solution to the flask, and swirl to mix well. Label this flask and keep it tightly capped until ready for use. 

Rinse the inside of a CLEAN buret three times with small quantities of your 0.1 M sodium hydroxide 
solution (called "rinsing in" with the solution to be used in the buret). Drain the rinses though the 
stopcock and tip. Do not forget to rinse liquid through the tip, to replace water there. Fill the buret above 
the 0.0 mL mark with 0.1 M NaOH, and then drain it until the meniscus is slightly below 

NOTE: Do not waste the time it takes to set the starting level to exactly 0.0 mL. It is more efficient and 
more accurate to set the level between 1 and 2 mL and read the starting level precisely. 



www.ckl2.org 556 





c 



Tips on Technique 

• To read the buret accurately, hold a white card with a black stripe behind the buret, with the black 
stripe below the meniscus, and the meniscus itself in front of the white region above the black stripe 
(see illustration). The meniscus will appear black against the white card. Keeping your eye level 
with the meniscus, read the buret. 

• Remember to estimate one more digit than those marked on the scale. 

• Remember that the buret scale reads increasing volume downward, not upward. 



Tips on Technique 

• Record the starting level to 0.01 mL precision (as always, estimate one more digit than marks indicate) . 

• Titrate the solution of primary acid standard with the 0.1 M NaOH until faint (see figure) phenolph- 
thalein color appears and persists for 30 seconds. (Why might the color slowly disappear even after 
all acid is titrated?) Record the final buret reading to 0.01 mL precision. 

• Mix the solution in the titration flask thoroughly after each addition of titrant, to ensure complete 
reaction before adding more. 

• As you near the endpoint, wash the sides of the flask with distilled water to make sure that all 
delivered titrant is in solution. 



557 



www.ckl2.org 



When you see that you are within a drop or two of the endpoint, split drops to avoid overshooting 
the endpoint. 




Lab Procedure 

Perform three titrations. For each, calculate the molarity of your NaOH solution. (From the mass of acid, 
and its molecular weight, you can calculate the number of moles of acid, which is equal to the number of 
moles of base you delivered. The molarity is found from the number of moles and the volume.) 

When you have three values for the molarity of your NaOH solution, determine the average value. 

Table 18.8: Data Table for Part I 



Trial 1 



Trial 2 



Trial 3 



Initial Reading, NaOH mL 

buret 

Final Reading, NaOH mL 

buret 

Volume of NaOH added mL 

Grams of acid standard g 

Moles of acid standard mol 

Molarity of NaOH M 



mL 
mL 

mL 

g 
mol 

M 



mL 
mL 

mL 

g 
mol 

M 



Average molarity of NaOH = M 

Procedure for Part II 

Part 2. Finding the Molar Mass of an Unknown Acid 

• Obtain a sample of a solid unknown from your instructor. Record its ID code in your report. 

• Also, record the approximate amount of unknown to use in each titration, and the number of acid 
hydrogens per molecule. Your instructor will provide this information. 

• Weigh the suggested amount into a clean 125 - mL erlenmeyer flask, by taring the balance with the 
flask on the pan, and then adding the acid to the flask. Record the mass of the sample to the precision 
allowed by the balance. Add 30 to 40 mL of distilled water and swirl to dissolve your sample. Add 
three to four drops of phenolphthalein indicator solution and swirl to mix well. 

• Titrate your sample with your standardized NaOH solution until faint phenophthalein color persists 
for 30 seconds. *If your titration requires 10 to 25 mL of NaOH solution, carry out a second titration 
with an unknown sample of about the same mass. Otherwise, adjust the sample mass to bring the 
expected end-point volume to between 10 and 25 mL and do two more titrations. 



www.ckl2.org 



558 



• For each titration, compute the molar mass of the unknown acid, to the precision allowed by your 
data. Do three titrations and report the average molar mass for the solid acid. 

Data for Part II 

ID code of acid 



Number of Acidic Hydrogens in Acid 
Approximate mass of acid to be used 



Table 18.9: Data Table for Part II 



Trial 1 



Trial 2 



Trial 3 



Mass of unknown acid g 
sample 

Volume of NaOH used mL 

Mols of NaOH used moL 

Moles of acid present 1 mol 

Molar mass of acid 2 g/mol 

Molarity of NaOH M 



mL 
moL 
mol 
g/mol 

M 



mL 
moL 
mol 
g/mol 

M 



1 Moles of acid present 

2 Molar mass of acid 
Average Molar Mass 



moles NaOH 



number of H + ions per acid molecule 

2 Molar mass of acid = g ' a " li acu 

moles acid 



_g/mol 



18.4 Buffers 

Lesson Objectives 

• Define a buffer and give various examples of buffers. 

• Explain the effect of a strong acid on the pH of a weak acid/conjugate base buffer. 

• Explain the effect of a strong base on the pH of a weak base/conjugate acid buffer. 

Introduction 

There are many situations in which it is desirable to keep the pH of a solution close to a particular value 
even though quantities of acids and/or bases are added to the solution. Many organic and biochemical 
reactions require acids or bases in the reaction but if the pH goes too high or too low, the products will be 
destroyed. For these reactions, it is necessary to keep the pH within a very small range even while acids 
or bases are added to the reaction. Chemists use mixtures called buffers to keep the reaction solutions 
within the necessary pH range. Buffers are mixtures of chemicals that cause a solution to resist changes 
in pH. 

Buffers are very important to many biological reactions. Human blood is a substance whose function is 
very dependent on the function of buffers. Human blood must maintain an almost constant pH between 7.3 
and 7.5 . If a person's blood pH goes 0.2 outside the acceptable range, the person will become unconscious 
and the blood pH goes 0.4 outside the range, the person will die. The pH of human blood can change 
depending on foods we eat and the rate at which we inhale and exhale C02- Fortunately, the human blood 



559 



www.ckl2.org 



stream has buffers which are able to resist pH changes due to food intake and breathing rates. 

Buffers 

A buffer is a solution that maintains the pH level when small amounts of acid or base are added to the 
system. Buffer solutions contain either a weak acid and the conjugate base of the weak acid or a weak 
base and the conjugate acid of the weak base. A common procedure for producing a buffer in the lab is 
to make a solution of a weak acid and a salt of that weak acid. For example, you could make a buffer by 
making a solution containing acetic acid and sodium acetate. 

The common buffer mentioned above would contain acetic acid, CH3COOH, and the acetate ion, CHj,COO~ . 
This buffer can keep a solution in the range of a pH of 3.7 - 5.8 even though small amounts of acids or 
bases are added to the solution. Another example is hydrogen phosphate ion, HPO^~, and the phosphate 
ion, PO^~ which will buffer a solution in the pH range of 11.3 - 13.3. 

How is it possible that a solution will not change its pH when an acid or base is added? Let's examine the 
acetic acid/acetate ion buffer. The ionization equation for acetic acid is shown below. 

HC2H S 2 ( cl q) £+ H + C 2 H 3 2 

It is obvious that if acid {H + ions) is added to this solution, the equilibrium will shift toward the reactants 
to use up some of the added hydrogen ions. Equilibrium will be re-established in the solution with different 
concentrations of the three species in the reaction. It is also obvious that if a base is added to this solution, 
the base will remove hydrogen ions and the equilibrium will shift to the right to partially counteract the 
stress. Again, equilibrium will be re-established with new concentrations. 

What is not obvious about this buffer solution is that acetic acid is a weak acid and therefore, most of 
the acid dissolved in the solution remain as acid molecules and do not dissociate. Therefore, there will a 
large quantity of undissociated HC 2 H^0 2 . The fact that we also dissolved sodium acetate in this solution 
provides a large quantity of acetate ions, CiH^O^ , in the solution. The existence of large quantities of both 
undissociated acid molecules and acetate ions in the solution is what allows the buffer to consume quite a 
large amount of added acid or base without the pH changing significantly. 

Examine what happens to 1.00 liter of pure water to which 0.100 mole of gaseous HCl is added. 

The original concentration of hydrogen ion in the pure water is 1.00 x 10~ 7 M and therefore, the pH is 7. 

After the 0.100 mole of HCl is added, the concentration of the hydrogen ion will be 0.100 M (plus the 
original 1.00 x 10~ 7 M which can be neglected as not significant). This new concentration of hydrogen ion 
will produce a pH = 1. So, the addition of the 0.100 mole of gaseous HCl caused the pH of the pure water 
to change from 7 to 1. 

Let's now see what happens if this same amount of gaseous HCl is added to an acetic acid-acetate ion 
buffer. We'll the same amount of original solution, 1.00 liter, and let's say we made this solution to contain 
0.50 M acetic acid and 0.50 M acetate ion (0.50 M sodium acetate which totally dissociated). The acetic 
acid dissociation equation will reach equilibrium and its K a will be 1.8 x 10 -5 . 

[HC 2 H 3 2 ] 

We know both the acetic acid and the acetate ion concentrations will be 0.50 M, so we can plug these 
values into the expression and solve for [H + ]. 

= (1.8X10-» 3 2] = (1.8 X 10^)(0.50) = 
1 ' [C2H3O5] (0.50) 

www.ckl2.org 560 



Then we can insert the hydrogen ion concentration into the pH formula and determine the original pH of 
the buffer solution. 

pH = -log(1.8 x 1(T 5 ) = 4.74 

Next, we will add the same 0.100 mole of gaseous HCl to this buffer solution and calculate the pH of the 
solution after the acid has been added and equilibrium has been re-established. 

When we add 0.100 mole HCl gas to this solution, the added hydrogen ion will combine with acetate ion 
to produce more undissociated acid. A small amount of the added hydrogen ions may remain as ions but 
the amount is, once again, beyond the significant figures of the problem and can be neglected. The new 
[HC 2 H 3 2 ] will equal 0.60 M (the original 0.50 M plus the added 0.10 M) and the new [C 2 H 3 2 ~] will 
equal 0.40M (the original 0.50 M minus the 0.10 M that reacted with the added hydrogen ions). We can 
now plug these values into the K a expression, calculate the new [H + ], and find the new pH. 

[H+][C 2 H 3 2 ] 

K a = -r-^ f- = 1.8 x 10~ 5 

[HC 2 H 3 2 ] 

r .. (1.8xl0- 5 )[//C 2 // 3 O 2 ] (1.8xl0- 5 )(0.60) 

\H + ] = - = J = - £ - = 2.7 x 10~ 5 M 

L J [C 2 H 3 0~] (0.40) 

pH = -log (2.7 x 10~ 5 ) = 4.57 

The same quantity of HCl gas that changed the pH of pure water from 7 to 1 has changed the pH of this 
buffer from 4.74 to 4.57. . . only a change of 0.17. . . that is the function of a buffer. Buffers resist change 
to pH. We could this same calculation but add a base instead of an acid and show that the pH increases 
by this same slight amount. Maybe that would be good practice for you. 

Sample question: Which of the following combinations would you expect possible to make into buffer 
solutions: 

(a) HCl0 4 /ClO A - 

(b) CH 3 NH 2 /CH 3 NH 3 + 
Solution: 

(a) HCIO4/CIO4 : HCIO4 is a strong acid and buffers are made from weak acids and their conjugate bases 
or weak bases and their conjugate acids. Therefore this cannot be made into a buffer solution. 

(b) CH 3 NH 2 /CH 3 NH 3 + : CH 3 NH 2 is a weak base and CH 3 NH 3 + is the conjugate acid of this base. Therefore 
this can be made into a buffer solution. 

The Buffer in Blood 

The primary buffer found in your bloodstream is carbonic acid, H 2 C0 3 . The carbonic acid is present due 
to carbon dioxide from your respiratory system dissolving in water. 

C0 2{g) + H 2 {L) i? H 2 C0 3{aq) Equation 1 

The amount of carbonic acid in your bloodstream is affected by the rate of your respiration. If you breathe 
rapidly, you reduce the amount of C0 2 in your bloodstream and the equilibrium shown in Equation 1 shifts 
toward the reactants thus lowering the amount of H 2 C0 3 . If you breathe slowly, the amount of C0 2 in 
your bloodstream increases and the equilibrium in Equation 1 shifts toward the products increasing the 
amount of H 2 C0 3 in your system. 

561 www.ckl2.org 



Once the H2CO3 is produced by dissolving carbon dioxide, the carbonic acid dissociates in your blood as 
shown in Equation 2. 

H2C0 3{aq) <=> H + + HC0 3 Equation 2 

This is the buffering reaction in your blood, composed of the weak acid, H2CO3, and its conjugate base, 
HC0 3 . 

Any changes in blood pH that could be caused by food intake, will be buffered by this equilibrium system. 
If acid is added to your blood, the equilibrium in Equation 2 will shift toward the reactants using up the 
hydrogen ions and if base were added to your blood thus reacting with hydrogen ions, the equilibrium will 
shift toward the products generating more hydrogen ions. This buffer in your blood is very efficient at 
keeping your blood pH in the necessary range. 

A problem exists for a few people because their respiratory rate is highly affected by their emotional 
state. Some people, when they get nervous, begin breathing very fast or very slow. Breathing too fast is 
called hyperventilating and breathing too slow is called hypoventilating. Your respiratory rate is normally 
controlled by the amount of carbon dioxide in the blood. Your body receives instructions to breathe faster 
or slower to adjust the amount of carbon dioxide in your blood in order to properly regulate the buffer 
system. When people people breathe too fast or too slow because of other reasons, the CO2 content of the 
blood becomes incorrect and the pH of the blood rises or lowers outside the acceptable range of 7.3 - 7.5 
. When this happens, the person passes out or becomes unconscious. People who hyperventilate when 
excited or nervous are sometimes advised to carry a lunch sack or something similar to breathe into when 
they are feeling light-headed. Breathing into a sack returns air with the same concentration of carbon 
dioxide that was exhaled . . . this keeps the amount of carbon dioxide in the blood up and avoids passing 
out. 

Lesson Summary 

• A buffer is a solution of a weak acid and its conjugate base or a weak base and its conjugate acid 
that resists changes in pH when an acid or base is added to it. 

• Adding a strong acid to a weak acid/conjugate base buffer only decreases the pH by a small amount. 

• Adding a strong base to a weak base/conjugate acid buffer only increases the pH by a small amount. 

Review Questions 

1. Define a buffer solution. 

2. What are two different types of buffer solutions? 

3. One of the following statements of buffers is incorrect. Which one? 

(a) A buffer may be prepared from a weak acid and its conjugate base salt. 

(b) A buffer may be prepared from a weak base and its conjugate acid salt. 

(c) A buffer is a solution that can resist changes in pH when any amount of acid or base is added 
to it. 

(d) A buffer is a solution that can resist changes in pH when a small amount of acid or base is 
added to it. 

4. Which pair of aqueous 1.0 mol/L solutions could be chosen to prepare a buffer? 

5. (a) i and iii only 

(b) i and iii only 

(c) i, ii and iii 

(d) None of these solutions is a buffer. 

www.ckl2.org 562 



i. NH^HSO^aq) and H 2 SO^ aq) 
h. HN0 2(aq) and NaN0 2{aq) 
hi. NH^Cl[a q ) and NH^ aq) 

6. Which of the following would form a buffer solution if combined in appropriate amounts? 

(a) HCl and NaCl 

(b) HCN and NaCAT 

(c) H 2 S and A^S 

(d) HN0 3 and AWV0 3 

7. A buffer is made up of a weak acid and a conjugate base. A small amount of acid is added to the 
buffer. What happens to the resulting solution? 

(a) The acid dissociation constant goes up. 

(b) The concentration of the weak acid in the buffer goes down. 

(c) The pH of the solution goes up. 

(d) The pH remains almost the same. 

8. Almonds from the wild have a very bitter taste because of hydrogen cyanide (and therefore are very 
dangerous to eat!!!). Interestingly if we think about HCN in a buffer situation, HCN and NaCN can 
be considered to act as a buffer solution. Sulfurous acid is used quite frequently as a cleansing agent. 
If we take the sodium salt, Na 2 SOs, of sulfurous acid, H 2 SOs we do not make a buffer solution. Why 
is this so? Why would one make a buffer solution and not the other? 

Further Reading / Supplemental Links 

• http://en.wikipedia.org 

Vocabulary 

buffer A buffer is a solution of a weak acid and its conjugate base or a weak base and its conjugate acid 
that resists changes in pH when an acid or base is added to it. 



Image Sources 



(i 

(2 
(3 
(4 
(5 
(6 
(7 
(8 
(9 



Hydrangeas. Public Domain. 
Therese Forsythe. . CC-BY-SA. 

Theresa Knott. Typical Titration Setup.. GNU Free Documentation License. 
Datamax. An electronic pH meter. Public Domain. 
Andel Fruh. Structure of Phenolphthalein.. Public Domain. 

Structure of Methyl Orange (Yellow Form).. Public Domain. 
Therese Forsythe. Titration Curve for a Strong Acid vs. Weak Base.. CC-BY-SA. 
Therese Forsythe. Titration Curve for a Strong Acid vs. Strong Base.. CC-BY-SA. 
Richard Parsons. The color of Universal Indicator at various pH's.. CC-BY-SA. 

563 www.ckl2.org 



Chapter 19 

Radioactivity and the Nucleus 



19.1 Discovery of Radioactivity 

Lesson Objectives 

The student will: 

• describe the roles played by Henri Becquerel and Marie Curie played in the discovery of radioactivity. 

• list the most common emissions from naturally radioactive nuclei. 

Introduction 

No one could have known in the 1800s that the discovery of the fascinating science and art form of 
photography would eventually lead to the splitting of the atom. The basis of photography is the fact 
that visible light causes certain chemical reactions. If the chemicals are spread thinly on a surface but 
protected from light by a covering, no reaction occurs. When the covering is removed, however, light 
acting on the chemicals causes them to darken. With millions of cameras in use today we do not think 
of it as a strange phenomenon, but at the time of its discovery photography was a strange and wonderful 
thing. Even stranger was the discovery by Roentgen, that radiation other than visible light could expose 
photographic film. He found that film wrapped in dark paper would react when x-rays went through the 
paper and struck the film. 

Becquerel and Radioactivity 

When Becquerel heard about Roentgen's discovery, he wondered if his fluorescent minerals would give 
the same x-rays. Becquerel placed some of his rock crystals on top of a well-covered photographic plate 
and sat them in the sunlight. The sunlight made the crystals glow with a bright fluorescent light, but 
when Becquerel developed the film he was very disappointed. He found that only one of his minerals, a 
uranium salt, had fogged the photographic plate. He decided to try again, and this time, to leave them 
out in the sun for a longer period of time. Fortunately, the weather didn't cooperate and Becquerel had to 
leave the crystals and film stored in a drawer for several cloudy days. Before continuing his experiments, 
Becquerel decided to check one of the photographic plates to make sure the chemicals were still good. To 
his amazement, he found that the plate had been exposed in spots where it had been near the uranium 
containing rocks and some of these rocks had not been exposed to sunlight at all. 

www.ckl2.org 564 



Bequerel's radioactive rock sitting on an envelope containing photographic film. 





Rock resting on covered film. 



Shadow on developed film. 



In later experiments, Becquerel confirmed that the radiation from the uranium had no connection with 
light or fluorescence, but the amount of radiation was directly proportional to the concentration of uranium 
in the rock. Becqueral had discovered radioactivity. 



The Curies and Radium 



One of Becquerel's assistants, a young Polish scientist named Maria Sklowdowska (to become Marie Curie 
after she married Pierre Curie), became interested in the phenomenon of radioactivity. With her husband, 
she decided to find out if chemicals other than uranium were radioactive. The Austrian government was 
happy to send the Curies a ton of pitchblende from the mining region of Joachimstahl because it was waste 
material that had to be disposed of anyway. The Curies wanted the pitchblend because it was the residue of 
uranium mining. From the ton of pitchblend, the Curies separated 0.10 g of a previously unknown element, 
radium, in the form of the compound, radium chloride. This radium was many times more radioactive 
than uranium. 

By 1902, the world was aware of a new phenomenon called radioactivity and of new elements which 
exhibited natural radioactivity. For this work, Becquerel and the Curies shared the 1903 Nobel Prize in 
Physics. For her subsequent work in radioactivity, Marie Curie won the 1911 Nobel Prize in Chemistry. 
She was the first female Nobel laureate and the only person ever to receive two Nobel Prizes in two different 
scientific categories. 

Marie Sklodowska before she moved to Paris. (Source: http : //chemistry . about . com/od/historyof chemistr 
ig/Pictures-of-Famous-Chemists/Marie. --bs.htm. Public Domain) 



565 



www.ckl2.org 




Further experiments provided information about the characteristics of the penetrating emissions from 
radioactive substances. It was soon discovered that there were three common types of radioactive emissions. 
Some of the radiation could pass easily through aluminum foil while some of the radiation was stopped by 
the foil. Some of the radiation could even pass through foil up to a centimeter thick. The three basic types 
of radiation were named alpha, beta, and gamma radiation. The actual composition of the three types of 
radiation was still not known. 

Eventually, scientists were able to demonstrate experimentally that the alpha particle, a, was a helium 
nucleus (a particle containing two protons and two neutrons), a beta particle, /?, was a high speed electron, 
and gamma rays, y, were a very high energy form of light (even higher energy than x-rays). 



Lesson Summary 

• Henri Becquerral, Marie Curie, and Pierre Curie shared the discovery of radioactivity. 

• The most common emissions of radioactive elements were called alpha (a), beta (/?), and gamma (y) 
radiation. 



Vocabulary 

alpha particle An alpha particle is a helium-4 nucleus. 

beta particle A beta particle is a high speed electron, specifically an electron of nuclear origin. 

gamma ray Gamma radiation is the highest energy on the spectrum of electromagnetic radiation. 

Marie Curie Marie Curie was a physicist and chemist of Polish upbringing, and subsequently, French 
citizenship; a pioneer in the field of radioactivity and the only person to ever win two Nobel prizes 
in science. 



www.ckl2.org 566 



Further Reading / Supplementary Links 

http : //www. chem . duke . edu/~ jds/cruise_chem/nuclear/discovery . html 

Review Questions 

1. Put the letter of the matching phrase on the line preceding the number. 1. alpha particle - a. 

high energy electromagnetic radiation 

2. beta particle - b. a high speed electron 

3. gamma ray - c. a helium nucleus 

19.2 Nuclear Notation 

Lesson Objectives 

The student will: 

• state the information contained in the atomic number of a nucleus. 

• state the information contained in the mass number of a nucleus. 

• subtract the atomic number from the mass number to determine the number of neutrons in a nucleus. 

• read and write complete nuclear symbols (know the structure of the symbols and understand the 
information contained in them). 

Introduction 

Just as chemical formulas use symbols and chemical equations use symbols, nuclei are represented by 
symbols. The complete nuclear symbol contains the symbol for the element and numbers that relate to 
the number of protons and neutrons in that particular nucleus. 

Atomic and Mass Numbers 

The identity of an atom is determined by the number of protons in the nucleus of the atom. The number 
of protons in the nucleus of the atom is also known as the ATOMIC NUMBER. The atomic number is 
the smaller number appearing on the periodic table for each atom. The atomic number for hydrogen is 1. 
This means that if a nucleus has 1 proton in it, it is a hydrogen nucleus no matter how many neutrons it 
has. The mass number for a nucleus is the total number of protons and neutrons (nucleons) in the nucleus 
of an atom. Both the mass number and the atomic number for nuclei are always whole numbers because 
there are no fractions of nucleons. To find the number of neutrons in the nucleus, you would subtract the 
atomic number from the mass number. For example, if the atomic number for a nucleus was 8 and the 
mass number was 18, the nucleus would contain 8 protons (equal to the atomic number) and 10 neutrons 
(18 nucleons - 8 protons = 10 neutrons). 

The Complete Nuclear Symbol 

To write a complete nuclear symbol, the mass number is placed at the upper left (superscript) of the 
chemical symbol and the atomic number is placed at the lower left (subscript) of the symbol. The complete 
nuclear symbol for helium-4 is drawn below. 

567 www.ckl2.org 



The complete nuclear symbol for helium-4. (Source: Richard Parsons. CC-BY-SA) 

mass number — 




_ _ _ chemical symbol 

atomic number 

The following nuclear symbols are for a nickel nucleus with 31 neutrons and a uranium nucleus with 146 
neutrons. 

59 l\Ti 238 j j 

28 i>x 92 *-* 

In the nickel nucleus represented above, the atomic number 28 indicates the nucleus contains 28 protons, 
and therefore, it must contain 31 neutrons in order to have a mass number of 59. The uranium nucleus 
has 92 protons as do all uranium nuclei and this particular uranium nucleus has 146 neutrons. Another 
way of representing these nuclei would be Ni - 59 and U - 238. 

Dalton's original atomic theory stated that all atoms of an element were identical in every way, but it was 
later discovered by Thomson that atoms of an element were identical in every way except they could have 
different mass numbers. This mass difference results from nuclei of the same element having a different 
number of neutrons. In order to be the same element, nuclei must have the same number of protons, but 
they may have a different number of neutrons. Atoms with the same atomic number but a different mass 
number are called isotopes. 



1 
1 



H *h :h 



hydrogen- 1 hydrogen-2 hydrogen-3 

(protium) (deuterium) (tritium) 

Hydrogen, for example, has three isotopes, shown in the figure above. All three of hydrogen's isotopes 
must have one proton (to be hydrogen), but they have zero, one, or two neutrons in the nucleus. 

Originally, the names protium, deuterium, and tritium were suggested for the three isotopes of hydrogen, 
but about a year after the names were suggested it was felt that a new name indicated a new element, 
which these were not. Furthermore, naming compounds using these names would cause great difficulty. 
Nuclear scientists today use the names hydrogen-1, hydrogen-2, and hydrogen-3, but occasionally you 
will see or hear the other names. 



Lesson Summary 

• The complete nuclear symbol has the atomic number (number of protons) of the nucleus as a subscript 
at the lower left of the chemical symbol and the mass number (numberofprotons + neutrons) as a 
superscript at the upper left of the chemical symbol. 



Vocabulary 

atomic number The atomic number indicates the number of protons in the nucleus. 
www.ckl2.org 568 



mass number The mass number indicates the number of protons plus the number of neutrons in the 
nucleus. 

electron An electron is a fundamental sub-atomic particle that carries a negative charge. 

neutron A neutron is a sub-atomic particle with no electric charge and a mass slightly larger than a 
proton. 

proton A proton is a fundamental sub-atomic particle with a net positive charge. 

nucleus The nucleus of an atom if the very dense region, consisting of nucleons (proton and neutrons) 
at the center of an atom. 

nuclei Nuclei is the plural of nucleus. 

nucleon A nucleon is a constituent part (proton or neutron) of an atomic nucleus. 

nuclide A type of nucleus specified by its atomic number and mass number. 

Review Questions 

1. Write the complete nuclear symbol for a nucleus of chlorine that contains 17 protons and 20 neutrons. 

2. Write the complete nuclear symbol for a nucleus of oxygen that contains 8 protons and 10 neutrons. 

3. If a nucleus of uranium has a mass number of 238, how many neutrons does it contain? 

4. In the nuclear symbol for a beta particle, what is the atomic number? 

5. Is it possible for isotopes to be atoms of different elements? Explain why or why not. 

6. How many neutrons are present in a nucleus whose atomic number is one and whose mass number is 
one? 

7. Name the element of an isotope whose mass number is 206 and whose atomic number is 82. 

8. How many protons and how many neutrons are present in a nucleus of lithium- 7? 

9. What is the physical difference between at/- 235 atom and at/- 238 atom? 
10. What is the difference in the chemistry of a U — 235 atom and at/- 238 atom? 



19.3 Nuclear Force 

Lesson Objectives 

The student will: 



compare the energy released per gram of matter in nuclear reactions to that in chemical reactions, 
express the equation for calculating the change in mass during nuclear reactions that is converted 
into energy, 
express the relationship between nuclear stability and the nuclei's binding energy per nucleon ratio. 

569 www.ckl2.org 



Introduction 

There are only four forces in nature that produce all interactions between objects. Gravity affects all 
particles and only attracts, never repels. It is the weakest of the four forces but acts over great distances. 
The electromagnetic force acts between electric charges and magnetic fields and causes all physical and 
chemical processes to occur. It can also act at large distances. The weak nuclear force is limited to the 
atomic nucleus and causes unstable particles and nuclei to decay. The strong nuclear force is also limited 
to the nucleus and binds quarks into nucleons and nucleons into nuclei. 

Nuclear Force Overcomes Proton Repulsion 

A nucleus (with one exception, hydrogen-1) consists of some number of protons and neutrons pulled 
together in an extremely tiny volume. Since protons are positively charged and like charges repel, it is 
clear that protons cannot remain together in the nucleus unless there is a powerful force holding them there. 
The force which holds the nucleus together is generated by nuclear binding energy. We are concerned not 
with just the total amount of binding energy a nucleus possesses, but also with how many nucleons (protons 
and neutrons) are present for the binding energy to hold together. It is more illuminating for us to consider 
the amount of binding energy per nucleon that a nucleus contains. A nucleus with a large amount of 
binding energy per nucleon will be held together tightly and is referred to as stable. When there is too 
little binding energy per nucleon, the nucleus will be less stable and may disintegrate (come apart). When 
nuclei come apart, they come apart violently accompanied by a tremendous release of energy in the form 
of heat, light, and radiation. 

Mass Defect Becomes Binding Energy 

We know that a proton has a mass of 1.00728 Daltons (also known as atomic mass units) and a neutron 
has a mass of 1.00867 Daltons. A helium-4 nucleus consists of two protons and two neutrons. The mass 
of these separate protons and neutrons would be: 

2 x 1.00728 + 2 x 1.00867 = 4.03190 Daltons. 

The actual mass of a helium-4 nucleus is known to be 4.00150 Daltons. It would appear that some mass 
has been lost when the particles formed a nucleus. This difference between the sum of the masses of the 
individual nucleons and the mass of the corresponding nucleus always occurs and has been called mass 
defect. The mass defect in this case is 0.03040 Daltons. This mass, of course, is not lost but is converted 
into energy - binding energy. Albert Einstein first theorized that mass and energy could be converted into 
one another and he produced the equation with which to calculate the conversion; E = mc 2 , where E is 
energy in Joules, m is the mass in kilograms, and c is the speed of light, 3 x 10 8 meters/second. If we think 
about the conservation of mass and energy, we can account for everything. Nothing is lost and nothing 
appears from nowhere. The conversion of a very small amount of mass into energy produces an immense 
amount of energy, as you might guess from the size of the number you get when you square the speed of 
light. If 1.00 gram of mass were converted to energy in this manner, it would produce 9.0 x 10 13 Joules. 
This amount of energy would raise the temperature of 300 million liters of water from room temperature 
to boiling. 

More Binding Energy per Nucleon Produces Stability 

It is conventional to plot the binding energy per nucleon (total binding energy divided by the number of 
nucleons in the nucleus) versus the atomic mass of the nucleus. Such a graph is shown in Figure 19.1. The 

www.ckl2.org 5 TO 



position of greatest binding energy per nucleon is held by iron-56. Nuclei both larger and smaller than 
iron-56 have less binding energy per nucleon and are therefore, less stable. In the graph below, binding 
energy is measured in Mev (million electron volts). One million electron volts is equal to 1.6 x 10 -13 Joules. 
(This graph has been smoothed. The actual graph line zig zags up and down a little.) 



c , 


^Fe 




o ' 




23S,, 


u 




~— — — —^ iL 


u 






3 


y^ 




Z 


f—C 




^ 






u 






Q. 


NHe 




CI 






c 






u 






c 






IXI 






CI 






c 








2 




CD 


1 H 


^- 



Mass Number 



Figure 19.1: The graph of binding energy per nucleon for atoms between atomic number 1 and 92. 



Lesson Summary 

• The proton-proton repulsion in a nucleus is overcome by binding energy to hold the nucleus together. 

• The sum of the masses of the individual components of a nucleus is greater than the mass of the 
nucleus and the "lost" mass is called mass defect. 

• Much of the mass defect is converted into binding energy according the Einstein equation, E 

• The stability of a nucleus is determined by the amount of binding energy present for each nucleon 

• Nuclei with lower binding energy per nucleon may disintegrate. 

• The nucleus with the greatest binding energy per nucleon is iron-56. 



2 

mc . 



Vocabulary 

binding energy Binding energy is the amount of energy that holds a nucleus together, and therefore, 
also the amount of energy required to decompose a nucleus into its component nucleons. 

mass defect Mass defect is the difference between the sum of the masses of the nuclear components and 
the mass of the corresponding nucleus. Much of this lost mass is converted into binding energy. 



nucleon Nucleon is a collective name for neutrons and protons. 



Further Reading / Supplementary Links 

http://www.wisegeek.com/what-is-the-strong-nuclear-force.htm 

571 



www.ckl2.org 



Review Questions 

1. Iron-56 is a very stable nucleus while cobalt-60 is an unstable nucleus. Which nucleus would you 
expect to have more binding energy per nucleon? 

2. Calculate the mass defect and binding energy for a mole of carbon- 14 given the data below. 

The molar mass of carbon- 14 is 14.003241 g/mol. 

The molar mass of a proton is 1.007825 g/mol. 

The molar mass of a neutron is 1.008665 g/mol. 

Mass in kilograms is converted into energy in Joules by multiplying the mass times the speed of light 
squared, E = mc 2 . 

The speed of light is 3.00 x 10 8 m/s. 

kg • m 2 /s 2 = Joules 



19.4 Nuclear Disintegration 

Lesson Objectives 



The student will: 

• list some naturally occurring isotopes of elements that are radioactive. 

• describe the three most common emissions during natural nuclear decay. 

• express the changes in the atomic number and mass number of a radioactive nuclei when an alpha 
particle is emitted. 

• express the changes in the atomic number and mass number of a radioactive nuclei when a beta 
particle is emitted. 

• express the changes in the atomic number and mass number of a radioactive nuclei when a gamma 
ray is emitted. 

• express that protons and neutrons are not indivisible and are composed of particles called quarks. 

• express the number of quarks that make up a proton or neutron. 

Introduction 

Under certain conditions, less stable nuclei alter their structure to become more stable nuclei. Those 
unstable nuclei that are larger nuclei than iron-56 spontaneously disintegrate by ejecting particles. This 
process of decomposing to form a different nucleus is called radioactive decay. Many nuclei are radioactive; 
that is, they decompose by emitting particles and in doing so, become a different nucleus. All nuclei with 
84 or more protons are radioactive and many elements with less than 84 protons have both stable and 
unstable isotopes. 

Types of Radioactive Decay 

In natural radioactive decay, three common emissions occur. When these emissions were originally ob- 
served, scientists were unable to identify them as some already known particle and so named them alpha 
particles (a), beta particles (/?), and gamma rays (y). At some later time, alpha particles were 
identified as helium-4 nuclei, beta particles were identified as electrons, and gamma rays as a form of 

www.ckl2.org 572 



electromagnetic radiation like x-rays except much higher in energy and even more dangerous to living 
systems. 

Alpha Decay 

The nuclear disintegration process that emits alpha particles is called alpha decay An example of a nucleus 
that undergoes alpha decay is uranium-238. The alpha decay of U - 238 is 

In nuclear equations, it is required that the sum of the atomic numbers on the reactant side equal the sum 
of the atomic numbers on the product side and the same is true for the mass numbers on the two sides. 
In this equation, 

atomic number : 92 = 2 + 90 

mass number: 238 = 4 + 234. 

Therefore, the equation is balanced. 

Another alpha particle producer is thorium-230. 



230 , 




4„ 


226„ 


90 Th 


~* 


2 He + 


88 Ra 



Confirm that this equation is correctly balanced. 

Beta Decay 

Another common decay process is beta particle production, or beta decay. It may occur to you that we 
have a logically difficult situation here. Nuclei do not contain electrons and yet during beta decay, an 
electron is emitted from a nucleus. At the same time that the electron is being ejected from the nucleus, 
a neutron is becoming a proton. It is tempting to picture this as a neutron breaking into two pieces with 
the pieces being a proton and an electron. That would be convenient for simplicity, but unfortunately that 
is not what happens; more about this at the end of this section. 

In order to insert an electron into a nuclear equation and have the numbers add up properly, an atomic 
number and a mass number had to be assigned to an electron. The mass number assigned to an electron is 
zero (0) which is reasonable since the mass number is the number of protons plus neutrons and an electron 
contains no protons and no neutrons. The atomic number assigned to an electron is negative one (-1) 
because that allows a nuclear equation containing an electron to balance atomic numbers. Therefore, the 
nuclear symbol representing an electron (beta particle) is 

°« 

_i e or ,\P 

Thorium-234 is a nucleus that undergoes beta decay. Here is the nuclear equation for this beta decay. 

234 ml 234„ 

90 Th -* -I e + 91 Pa 

Note that both the mass numbers and the atomic numbers add up properly: 

573 www.ckl2.org 



atomic number: 90 = -1 + 91 

mass number: 234 = + 234 

The mass numbers of the original nucleus and the new nucleus are the same because a neutron has been 
lost, but a proton has been gained and so the sum of protons plus neutrons remains the same. The 
atomic number in the process has been increased by one since the new nucleus has one more proton than 
the original nucleus. In this beta decay, a thorium-234 nucleus has become a protactinium-234 nucleus. 
Protactinium-234 is also a beta emitter and produces uranium-234. 



234„ 









234 XT 


^ Pa 


— > 


„e 


+ 


™ u 


91 




-1 




92 



Once again, the atomic number increases by one and the mass number remains the same; confirm that the 
equation is correctly balanced. 

Protons and Neutrons are Made Up of Quarks 

Protons and neutrons are not fundamental particles as electrons are. Protons and neutrons are composed 
of more fundamental particles called quarks. Quarks are fundamental particles not similar to electrons, 
but like electrons in that they have no smaller pieces. (If you think "quarks" is a somewhat fanciful name, 
the world of particle physics holds a lot of surprises for you.) Quarks come in six different types (particle 
physicists call them flavors). The six "flavors" of quarks are named up, down, top, bottom, strange, and 
charmed. Protons and neutrons are composed of a combination of up and down quarks. The up quark 
carries a charge of +2/3 and the down quark carries a charge of -1/3. A proton is composed of two up 
quarks and one down quark. Thus the charge on a proton is +2/3 + 2/3 - 1/3 = 1. A neutron is composed 
of one up quark and two down quarks, hence its charge is +2/3 - 1/3 - 1/3 = 0. With this information, 
you can see that a neutron is NOT composed of a proton and an electron. The beta particle produced 
during beta decay is created in the process of a neutron decaying to a proton. We will view the process in 
terms of the net effect which is changing a neutron into a proton and emitting an electron. 

Gamma Radiation 

Frequently, gamma ray production accompanies nuclear reactions of all types. In the alpha decay of £7-238, 
two gamma rays of different energies are emitted in addition to the alpha particle. 

238 TT 4 TT 234^, n 

92 U ^2 He+ 90 Th+ V 

Virtually all of the nuclear reactions in this chapter also emit gamma rays, but for simplicity the gamma 
rays are generally not shown. Nuclear reactions produce a great deal more energy than chemical reactions. 
Chemical reactions release the difference between the chemical bond energy of the reactants and products, 
and the energies released have an order of magnitude of 1 x 10 3 kJ/mol. Nuclear reactions release some of 
the binding energy and may convert tiny amounts of matter into energy. The energy released in a nuclear 
reaction has an order of magnitude of 1 x 10 8 kJ/mol. 

Decay Series 

The decay of a radioactive nucleus is a move toward becoming stable. Often, a radioactive nucleus cannot 
reach a stable state through a single decay. In such cases, a series of decays will occur until a stable nucleus 

www.ckl2.org 574 



is formed. The decay of U — 238 is an example of this (Table 19.1). The U - 238 decay series starts with 

U — 238 and goes through fourteen separate decays to finally reach a stable nucleus, Pb — 206. There are 

similar decay series for U — 235 and Th — 232. The U — 235 series ends with Pb — 207 and the Th — 232 series 
ends with Pb - 208. 

Table 19.1: 

Nucleus Type of Decay Product 

U - 238 a Th- 234 

Th - 234 B Pa- 234 

Pa - 234 B U - 234 

U - 234 a Th- 230 

Th - 230 a Ra - 226 

Ra - 226 a Rn - 222 

Rn - 222 a Po - 218 

Po- 218 a Pb- 214 

P»-214 /? Bi -214 

5/ -214 /? Po-214 

Po-214 a P6-210 

P6-210 /? Bi -210 

5/ -210 /? Po-210 

Po - 210 a- Pb - 206 



Several of the radioactive nuclei that are found in nature are present there because they are produced in 
one of the radioactive decay series. That is to say, there may have been radon on the earth at the time of 
its formation, but that original radon would have all decayed by this time. The radon that is present now 
is present because it was formed in a decay series. 

Lesson Summary 

• A nuclear reaction is one that changes the structure of the nucleus of an atom. 

• The atomic numbers and mass numbers in a nuclear equation must be balanced. 

• Protons and neutrons are made up of quarks. 

• The two most common modes of natural radioactivity are alpha decay and beta decay. 

• Most nuclear reactions emit energy in the form of gamma rays. 

Vocabulary 

alpha decay Alpha decay is a common mode of radioactive decay in which a nucleus emits an alpha 
particle (a helium-4 nucleus). 

beta decay Beta decay is a common mode of radioactive decay in which a nucleus emits beta particles. 
The daughter nucleus will have a higher atomic number than the original nucleus. 

quark Quarks are physical particles that form one of the two basic constituents of matter. Various species 
of quarks combine in specific ways to form protons and neutrons, in each case taking exactly three 
quarks to make the composite particle. 

5T5 www.ckl2.org 



Further Reading / Supplementary Links 

Video showing nuclear reactions. 

http : //www. youtube . com/watch?v=_f 4kYYAQC3c 

Short animation showing the bending of a, ft, and y radiation under the influence of electrically charged 
plates. 

http : //www.mhhe . com/physsci/chemistry/chemistryessential/f lash/radioa7 . swf 

Review Questions 

1. Write the nuclear equation for the alpha decay of radon- 198. 

2. Write the nuclear equation for the beta decay of uranium-237. 

3. There are six known quarks. The experimenter who discovers particles in nuclear physics gets the 
right to name the new particle. This has resulted in some very fanciful names for quarks. The six 
quarks are named, up quarks, down quarks, charmed quarks, strange quarks, top quarks, and 
bottom quarks. (The top and bottom quarks were originally named truth and beauty quarks, but 
the names were changed for some reason.) Protons and neutrons are each made of only up and 
down quarks and they are made of three quarks each. The up quark carries a charge of +o and the 
down quark carries a charge of — g. Determine by the final charge on the proton and neutron, what 
combination of three up and down quarks are required to make a proton and what combination will 
make a neutron? 

19.5 Nuclear Equations 

Lesson Objectives 

The student will: 

• define and give examples of fission and fusion. 

• classify nuclear reactions as fission or fusion. 

• correctly fill in the missing particle in a nuclear equation with one species missing 

• write balanced equations for nuclear transmutations. 

Introduction 

Atomic nuclei with an inadequate amount of binding energy per particle are unstable and occasionally 
disintegrate in an organized fashion. Such disintegrations are referred to as natural radioactivity. It is also 
possible for scientists to smash nuclear particles together and cause nuclear reactions between normally 
stable nuclei. These disintegrations are referred to artificial radioactivity. None of the elements above 
#92 on the periodic table occur on earth naturally . . . they are all products of artificial radioactivity 
(man-made). 

Fission and Chain Reactions 

Nuclei that are larger than iron-56 become smaller and in the process become more stable. These large 
nuclei undergo nuclear reactions in which they break up into two or more smaller nuclei. These reactions 
are called fission reactions. 

www.ckl2.org 576 



Conversely, nuclei that are smaller than iron-56 become larger nuclei in order to be more stable. These 
nuclei undergo a nuclear reaction in which smaller nuclei join together to form a larger nucleus. Such 
nuclear reactions are called fusion reactions. In both fission and fusion, large amounts of energy are given 
off in the form of heat, light, and gamma radiation. Nuclear fission was discovered in the late 1930s when 
U - 235 nuclides were bombarded with neutrons and were observed to split into two smaller-mass nuclei. 



1 _L 235 TT 

n + 92 U 



141 92 1 

56 Ba + 36 Kr + 3 n 



The products shown are only one of many sets of products from the disintegration of a U — 235 nucleus. 
Over 35 different elements have been observed in the fission products of U — 235. 




> fc 




n n 




When a U — 235 nucleus captures a neutron, it undergoes fission producing two lighter nuclei and three 
free neutrons (Figure ??). The production of the free neutrons makes it possible to have a self-sustaining 
fission process - a nuclear chain reaction. If at least one of the neutrons goes on to cause another 
U - 235 disintegration, the fission will be self-sustaining. If none of the neutrons goes on to cause another 
disintegration, the process dies out. If the mass of fissionable material is too small, the neutrons escape 
from the mass without causing another reaction and the reaction is said to be subcritical. When the 
mass of fissionable material is large enough, at least one of the neutrons will cause another reaction and 
the process will continue at a steady rate. This process is said to be critical. 

When the critical mass is reached, the neutron will always cause another disintegration. 



577 



www.ckl2.org 




Rod x 

When the mass is small enough, the neutron 
may escape without causing another reaction. 
When the mass is large, however, the neutron is 
very unlikely to escape. 

If enough mass is present for the reaction to escalate rapidly, the heat buildup causes a violent explosion. 
This situation is called supercritical. The amount of mass necessary to maintain a chain reaction differs 
with each fissionable material and is called that material's critical mass. 

A short animation of nuclear fission can be viewed at http://www.classzone.com/cz/books/woc_ 
07/resources/htmls/ani_chem/chem_f lash/popup. html?layer=act&src=qtiwf_act 129. 1 .xml 

Natural and Artificial Radioactivity 

There are several ways in which nuclei can undergo reaction and change their identity. Some nuclei are 
unstable and spontaneously emit particles and electromagnetic radiation. Such spontaneous disintegration 
is known as natural radioactivity. 

It is also possible to cause nuclear disintegration by striking a nucleus with another particle. This is 
called artificial radioactivity. Ernest Rutherford produced the first induced or artificial transmutation 
of elements by bombarding a sample of nitrogen gas with alpha particles. 



238 TT 1 


239 TT 


239_ 


239„ 


92 U + n - 


^ U - 
92 


--i e +93 Np - 


--I e +94 PU 



Irene Joliet-Curie, the daughter of Pierre and Marie Curie, carried out a transmutation that produced 
an unstable (radioactive) nucleus from a stable nucleus. She bombarded aluminum with alpha particles, 
and after the bombardment was stopped the product continued to emit radiation. It was determined 
that the alpha particle bombardment transmuted aluminum-27 nuclei to phosphorus-30 nuclei. The 
phosphorus-30 is a positron emitter. Positrons are subatomic particles with the same mass as an electron, 
but carry a positive one (+1) charge instead of negative one (-1). 

4 TT 27 _, 30„ 1 

2 He + 13 A1 -> l5 P + Q n 

30 30 

15 P "» 14 Sl + 1 P 

Fusion 

The nuclear reaction shown below is a reaction in which a lithium-7 nucleus combines with a hydrogen-1 
nucleus and produces two helium-4 nuclei and a considerable amount of energy. 

www.ckl2.org 578 



17 4 

1 H + Li -> 2 2 He + energy 

The combined mass of the lithium and hydrogen nuclei is 8.02329 Daltons and the combined mass of the 
two helium nuclei is 8.00520 Daltons. The mass lost is 0.01809 Daltons and is accounted for in the energy 
that is released. Nuclear reactions, in which two or more lighter-mass nuclei join together to form a single 
nucleus, are called fusion reactions or nuclear fusions. Of particular interest are fusion reactions in which 
hydrogen nuclei combine to form helium. Hydrogen nuclei are positively charged and repel each other. 
The closer the particles come, the greater is the force of repulsion. In order for fusion reactions to occur, 
the hydrogen nuclei must have extremely high kinetic energies so the velocities can overcome the forces 
of repulsion. These kinetic energies only occur at extreme temperatures such as those that occur in the 
cores of the sun and other stars. Nuclear fusion is the power source for the stars where the necessary 
temperature to ignite the fusion reaction is provided by massive gravitational pressure. 

The energy that comes from the sun and other stars is produced by fusion. (Source: http: //commons. 
wikimedia.org/wiki/File: Sun- in-X-ray. Public Domain) 




The heat to ignite the fusion reaction in thermonuclear weapons (hydrogen bombs) is provided by a small 
fission reaction that is set off inside the mass of hydrogen. It is hoped that the fusion reactions that will 
someday occur in fusion reactors will be ignited by multiple lasers. 

The conversion of hydrogen to helium in the sun requires several steps, but the net result is the fusion of 
four hydrogen-1 nuclei into one helium-4 nucleus. 

14 

4 1 H -> 2 He + 2 1 p + energy 

In stars more massive than our sun, fusion reactions involving carbon and nitrogen are possible. These 
reactions produce more energy than hydrogen fusion reactions. 

The exact reactions involved in thermonuclear weapons are secret, but they most likely involve either a 
reaction between two hydrogen-2 nuclei or a reaction between a hydrogen-2 nucleus and a hydrogen-3 
nucleus. 

2 2 4 

H + H -» He + energy 

2 XT 3 XT 4 XT 1 

jH + H -> 2 He + Q n + energy 

579 www.ckl2.org 



Intensive research is now being conducted to develop fusion reactors for electricity generation. The two 
major problems slowing up the development is finding a practical means for generating the intense tem- 
perature needed and developing a container than won't melt under the conditions of a fusion reaction. 
Electricity-producing fusion reactors are still a distant dream. 

A short animation of nuclear fusion can be viewed at http://www.classzone.com/cz/books/woc_ 
07/resources/htmls/ani_chem/chem_f lash/popup. html?layer=act&src=qtiwf _act 130. 1 .xml 

Lesson Summary 

• In nuclear reactions, the sum of the atomic numbers and the sum of the mass numbers on the two 
sides of the equation must be equal. 

• Naturally radioactive elements exist in the earth and are either alpha or beta emitters. 

• Artificial transmutation of elements can be accomplished by bombarding the nuclei of some elements 
with alpha or subatomic particles. 

• Nuclear fission refers to the splitting of atomic nuclei. 

• Nuclear fusion refers to the joining together to two or more smaller nuclei to form a single nucleus. 

Vocabulary 

artificial radioactivity Induced radioactivity that is produced by bombarding an element with high- 
velocity particles. 

chain reaction A multi-stage nuclear reaction that sustains itself in a series of fissions in which the 
release of neutrons from the splitting of one atom leads to the splitting of others. 

critical mass The smallest mass of a fissionable material that will sustain a nuclear chain reaction at a 
constant level. 

fission A nuclear reaction in which a heavy nucleus splits into two or more smaller fragments, releasing 
large amounts of energy. 

fusion A nuclear reaction in which nuclei combine to form more massive nuclei with the simultaneous 
release of energy. 

natural radioactivity The radioactivity that occurs naturally, as opposed to induced radioactivity. Also 
known as spontaneous fission. 

Further Reading / Supplementary Links 

http : //www. atomicarchive . com/Movies/index_movies . shtml 

Review Questions 

1. Only one particle is missing from this, equation. What are its atomic and mass numbers? 

7 N + 2 He -» .H + ? 

2. To what element does the missing particle in question #1 belong? 
www.ckl2.org 580 



3. When a U — 235 nucleus is struck by a neutron, the nucleus may be split into Ce — 144 and Sr — 90 
nuclei, also emitting four electrons and two neutrons. Write the equation for this nuclear reaction. 

4. Complete the following nuclear equatioqljg supplying the missing particle. 



85 



At 



5. Complete the following nuclear equationjjjg supplying the missing particle. 



84 



Po 



6. Complete the following nuclear equation Jgr supplying the missing particle. 



86 



Rn 



7. Complete the following nuclear equatioj^W supplying ^j^jmissing particle. 

80 kg+f ^ 79 AU 

8. Complete the following nuclear equation by supgtygng the pissing particle. 

? + "» 84 Po + -i e 

9. 



c j 


^Fe 




o ' 




238, . 


d) 




— — — _u 


o 






2 


J 1 ^ 




z 


1T"~C 




k. 






> 


K 4 He 




CJI 






k_ 






01 






C 






HI 






ra 






£ 






E 


2 




55 


1 H 





Mass Number 



Use information in the chart above to decide if carbon- 12 nuclei were to be transmuted into other nuclei 
that were more stable, would this more likely be accomplished by fission or by fusion? 

19.6 Radiation Around Us 

Lesson Objectives 

The student will: 

• compare qualitatively the ionizing and penetration power of a, fi, and y particles. 

• calculate the amount of radioactive material that will remain after an integral number of half-lives. 

• describe how carbon-14 is used to determine the age of carbon containing objects. 

Introduction 



All of us are subjected to a certain amount of ionizing radiation every day. This radiation is called 
background radiation and comes from a variety of natural and artificial radiation sources. Approximately 
82% of background radiation comes from natural sources. These include 1) sources in the earth; naturally 
occurring radioactive elements which are incorporated in building materials and also in the human body, 
2) sources from space in the form of cosmic rays, and 3) sources in the atmosphere such as radioactive 
radon gas released from the earth and radioactive atoms like carbon-14 produced in the atmosphere by 
bombardment from high-energy cosmic rays. 



581 



www.ckl2.org 



Background Radiation 

Approximately 15% of background radiation comes from medical x-rays and nuclear medicine. The re- 
maining 3% of background radiation comes from man-made sources such as: smoke detectors, luminous 
dials and signs, radioactive contamination due to historical nuclear weapons testing, normal operation of 
facilities used for nuclear power and scientific research, emissions from burning fossil fuels (primarily coal- 
burning power plants without ash-capture facilities), emissions from the improper disposal of radioactive 
materials used in nuclear medicine 

Public health agencies do not feel that the level of background radiation is a serious threat to public health, 
but they recommend that individuals limit their exposure to ionizing radiation as much as possible . To 
this goal, the medical profession has significantly reduced the number of x-rays recommended: skin tests for 
tuberculosis are recommended over x-rays and most dentists recommend dental x-rays every other check-up 
instead of every check-up. 

The Ionizing and Penetration Power of Radiation 

With all the radiation from natural and man-made sources, we should quite reasonably be concerned about 
how all the radiation might affect our health. The damage to living systems is done by radioactive emissions 
when the particles or rays strike tissue, cells, or molecules and alter them. These interactions can alter 
molecular structure and function; cells no longer carry out their proper function and molecules, such as 
DNA, no longer carry the appropriate information. Large amounts of radiation are very dangerous, even 
deadly. The ability of radiation to damage molecules is analyzed in terms of what is called ionizing power. 
When a radiation particle interacts with atoms, the interaction can cause the atom to lose electrons and 
thus become ionized. The greater the likelihood that damage will occur by an interaction is the ionizing 
power of the radiation. Much of the threat from radiation is involved with the ease or difficulty of protecting 
oneself from the particles. How thick of a wall do you need to hide behind to be safe? The ability of each 
type of radiation to pass through matter is expressed in terms of penetration power. The more material 
the radiation can pass through, the greater the penetration power and the more dangerous they are. 

Comparing only the three common types of ionizing radiation, alpha particles have the greatest mass. 
Alpha particles have approximately four times the mass of a proton or neutron and approximately 8, 000 
times the mass of a beta particle. Because of the large mass of the alpha particle, it has the highest ionizing 
power and the greatest ability to damage tissue. That same large size of alpha particles, however, makes 
them less able to penetrate matter. They collide with molecules very quickly when striking matter, add 
two electrons and become a harmless helium atom. Alpha particles have the least penetration power and 
can be stopped by a thick sheet of paper. They are also stopped by the outer layer of dead skin on people. 
This may seem to remove the threat from alpha particles but only from external sources. In a situation 
like a nuclear explosion or some sort of nuclear accident where radioactive emitters are spread around in 
the environment, the emitters can be inhaled or taken in with food or water and once the alpha emitter is 
inside you, you have no protection at all. 

Beta particles are much smaller than alpha particles and therefore, have much less ionizing power (less 
ability to damage tissue), but their small size gives them much greater penetration power. Most resources 
say that beta particles can be stopped by a one-quarter inch thick sheet of aluminum. Once again, however, 
the greatest danger occurs when the beta emitting source gets inside of you. 

Gamma rays are not particles but a high energy form of electromagnetic radiation (like x-rays except more 
powerful). Gamma rays are energy that has no mass or charge. Gamma rays have tremendous penetration 
power and require several inches of dense material (like lead) to shield them. Gamma rays may pass all 
the way through a human body without striking anything. They are considered to have the least ionizing 
power and the greatest penetration power. 

www.ckl2.org 582 



When researching thicknesses and materials required to stop various types of radiation, different estimates 
are encountered. There are apparently two reasons for this: 1) some estimates are based on stopping 95% 
of beta particles while other estimates are based on stopping 99% of beta particles, and 2) different beta 
particles emitted during nuclear reactions may have very different energies - some very high energy beta 
particles may not be stopped by the normal barrier. The safest amount of radiation to avoid damage to 
the human body is zero. It isn't possible to be exposed to zero ionizing radiation so the next best goal is 
to be exposed to as little as possible. 



Definition of Half-Life 

During natural radioactive decay, not all atoms of an element are instantaneously changed to atoms of 
another element. The decay process takes time and there is value in being able to express the rate at which 
a process occurs. In chemical reactions as well as radioactive decay, a useful concept is half-life, which 
is the time required for half of the starting material to be consumed. Half-lives can be calculated from 
measurements on the change in mass of a nuclide and the time it takes to occur. For a particular group of 
radioactive nuclei, it is not possible to know which nuclei will disintegrate or when they will disintegrate. 
The only thing we know is that in the time of that substance's half-life, half of the original nuclei will 
disintegrate. 

Selected Half-Lives 

The half-lives of many radioactive isotopes have been determined and they have been found to range from 
extremely long half-lives of 10 billion years to extremely short half-lives of fractions of a second (Table 
19.2). 

Table 19.2: TABLE OF SELECTED HALF-LIVES 



ELEMENT 


MASS 
BER 


NUM- 


HALF-LIFE 


ELEMENT 


MASS 
BER 


NUM- 


HALF-LIFE 


Uranium 


238 




4.5 billion 
years 


Californium 


251 




800 years 


Neptunium 


240 




1 hour 


Nobelium 


254 




3 seconds 


Plutonium 


243 




5 hours 


Carbon 


14 




5, 770 years 


Americium 


246 




25 minutes 


Carbon 


16 




0.74 seconds 



The quantity of radioactive nuclei at any given time will decrease to half as much in one half-life. For 
example, if there were 100 g of Cf - 251 in a sample at some time, after 800 years, there would be 50 g of 
Cf- 251 remaining and after another 800 years, there would only be 25 g remaining. 

Radioactive Dating 

An ingenious application of half-life studies established a new science of determining ages of materials by 
half-life calculations. For geological dating, the decay of U — 238 can be used. The half-life of U — 238 
is 4.5 x 10 9 years. The end product of the decay of U — 238 is Pb — 206. After one half-life, a 1.00 gram 
sample of uranium will have decayed to 0.50 grams of U — 238 and 0.43 grams of Pb - 206. By comparing 
the amount of U - 238 to the amount of Pb - 206 in a sample of uranium mineral, the age of the mineral 
can be estimated. Present day estimates for the age of the Earth's crust from this method is at least 4 
billion years. 



583 



www.ckl2.org 



Organic material is radioactively dated using the long-lived nuclide of carbon, carbon- 14. This method 
of determining the age of organic material was given the name radiocarbon dating. The carbon dioxide 
consumed by living systems contains a certain concentration of 14 C02- When the organism dies, the 
acquisition of carbon-14 stops but the decay of the C — 14 in the body continues. As time goes by, the 
ratio of C — 14 to C — 12 decreases at a rate determined by the half-life of C — 14. Using half-life equations, 
the time since the organism died can be calculated. These procedures have been used to determine the age 
of organic artifacts and determine, for instance, whether art works are real or fake. 

Lesson Summary 

• 82% of background radiation comes from natural sources. These include 1) sources in the earth; 
naturally occurring radioactive elements which are incorporated in building materials and also in the 
human body, 2) sources from space in the form of cosmic rays, and 3) sources in the atmosphere such 
as radioactive radon gas released from the earth and radioactive atoms like carbon-14 produced in 
the atmosphere by bombardment from high-energy cosmic rays. 

• Approximately 15% of background radiation comes from medical x-rays and nuclear medicine. The 
remaining 3% of background radiation comes from man-made sources such as: smoke detectors, 
luminous dials and signs, radioactive contamination due to historical nuclear weapons testing, normal 
operation of facilities used for nuclear power and scientific research, emissions from burning fossil fuels 
(primarily coal-burning power plants without ash-capture facilities), emissions from the improper 
disposal of radioactive materials used in nuclear medicine. 

• Of the three common nuclear emissions, alpha particles produce the greatest damage to cells and 
molecules but are the least penetrating. Gamma rays are the most penetrating but generated the 
least damage. 

• C — 14 dating procedures have been used to determine the age of organic artifacts. 

Vocabulary 

background radiation Radiation that comes from environment sources including the earth's crust, the 
atmosphere, cosmic rays, and radioisotopes. These natural sources of radiation account for the largest 
amount of radiation received by most people. 

half-life The half-life of a radioactive substance is the time interval required for a quantity of material 
to decay to half its original value. 

Further Reading / Supplementary Links 
Review Questions 

1 . Which of the three common emissions from radioactive sources requires the heaviest shielding? 

2. The half-life of radium-226 is about 1600 years. How many grams of a 2.00 gram sample will remain 
after 4800 years? 

3. Sodium-24 has a half-life of about 15 hours. How much of an 16.0 grams sample of sodium-24 will 
remain after 60.0 hours? 

4. A radioactive isotope decayed from 24.0 grams to 0.75 grams in 40.0 years. What is the half-life of 
the isotope? 

5. What nuclide is commonly used in the dating of organic artifacts? 

6. Why does an ancient wood artifact contain less carbon-14 than a piece of lumber sold today? 

www.ckl2.org 584 



7. The half-life of C — 14 is about 5, 700 years. An organic relic is found to contain C — 14 and C — 12 
in a ratio that is about one-eighth as great as the ratio in the atmosphere. What is the approximate 
age of the relic? 

8. Even though gamma rays are much more penetrating than alpha particles, it is the alpha particles 
that are more likely to cause damage to an organism. Explain why this is true. 

9. The radioactive isotope calcium-47 has been used in the study of bone metabolism; radioactive 
iron-59 has been used in the study of red blood cell function; iodine-131 has been used in both 
diagnosis and treatment of thyroid problems. Suggest a reason why these particular elements were 
chosen for use with the particular body function. 

19.7 Applications of Nuclear Energy 

Lesson Objectives 

The student will: 

• trace the energy transfers that occur in a nuclear power reactor from the binding energy of the nuclei 
to the electricity that leaves the plant. 

• define the term "breeder reactor." 

• list some medical uses of nuclear energy. 

Introduction 

It is unfortunate that when the topics of radioactivity and nuclear energy come up, most thoughts probably 
go to weapons of war. The second thought might be about the possibility of nuclear energy contributing to 
the solution of the energy crisis. Nuclear energy, however, has many applications beyond bombs and the 
generation of electricity. Radioactivity has huge applications in scientific research, several fields of medicine 
both in terms of imaging and in terms of treatment, industrial processes, some very useful appliances, and 
even in agriculture. 

Fission Reactors 

A nuclear reactor is a device in which a nuclear chain reaction is carried out at a controlled rate. When 
the controlled chain reaction is a fission reaction, the reactor is called a fission reactor. Fission reactors are 
used primarily for the production of electricity although there are a few fission reactors used for military 
purposes and for research. The great majority of electrical generating systems all follow a reasonably 
simple design. The electricity is produced by spinning a coil of wire inside a magnetic field. 




When the loop is spun, electric current is produced. The direction of the electric current is in one end of 
the loop and out the other end. The machine built to accomplish this task is called an electric generator 

585 www.ckl2.org 



(see figure above). Another machine usually involved in the production of electric current is a turbine. 
Although actual turbines can get very complicated, the basic idea is simple. Most of you have seen a child's 
pinwheel toy. When you blow through the fan blades of the pin wheel, it spins. A turbine is a pipe with 
many fan blades attached to an axle that runs through the pipe. 

When a fluid (air, steam, water) is forced through the pipe, it spins the fan blades which in turn spin the 
axle. To generate electricity, the axle of a turbine is attached to the loop of wire in a generator. When a 
fluid is forced through the turbine, the fan blades turn, the turbine axle turns, and the loop of wire inside 
the generator turns, thus generating electricity. 

A steam turbine. (Source: CK-12 Foundation. CC-BY-SA) 




The essential difference in various kinds of electrical generating systems is the method used to spin the 
turbine. For a wind generator, the turbine is a windmill. In a geothermal generator, steam from a geyser 
is forced through the turbine. In hydroelectric generating plants, water falling over a dam passes through 
the turbine and spins it. In fossil fuel (coal, oil, natural gas) generating plants, the fossil fuel is burned 
and the heat is used to boil water into steam and then the steam passes through the turbine and makes it 
spin. In a fission reactor generating plant, a fission reaction is used to boil the water into steam and the 
steam passes through the turbine to make it spin. Once the steam is generated by the fission reaction, a 
nuclear power plant is essentially the same as a fossil fuel plant. 

Naturally occurring uranium is composed almost totally of two uranium isotopes. It contains more than 
99% uranium-238 and less than 1% uranium-235. It is the uranium-235, however, that is fissionable 
(will undergo fission). In order for uranium to be used as fuel in a fission reactor, the percentage of 
uranium-235 must be increased, usually to about 3%. (Uranium in which the U — 235 content is more 
than 1% is called enriched uranium.) Somehow, the two isotopes must be separated so that enriched 
uranium is available for use as fuel. Separating the isotope by chemical means (chemical reactions) is 
not successful because the isotopes have exactly the same chemistry. Remember that chemical reactions 
are controlled by the electron configuration of the atom and all the isotopes of an element have the same 
electron configuration and hence the same chemistry. The only essential difference between U - 238 and 
U - 235 are their atomic masses and as a result the separation of the two isotopes will require a physical 
means that takes advantage of their difference in mass. The modern physical means used to separate the 
isotopes of uranium involve a gas centrifuge. Research separation of isotopes can also be accomplished 
with a mass spectrograph. You may wish to research these techniques on your own. 

Once the supply of U — 235 is acquired, it is placed in a series of long cylindrical tubes called fuel rods. 
These fuel cylinders are bundled together with control rods (Figure ??) made of neutron-absorbing 
material. The amount of U— 235 in all the fuel rods taken together is adequate to carry on a chain reaction 
but is less than the critical mass. (In the United States, all public nuclear power plants contain less than 
a critical mass of U — 235 and therefore, could never produce a nuclear explosion. The amount of heat 

www.ckl2.org 586 



generated by the chain reaction is controlled by the rate at which the nuclear reaction occurs. The rate of 
the nuclear reaction is dependent on how many neutrons are emitted by one U — 235 nuclear disintegration 
and strike a new U — 235 nucleus to cause another disintegration. The purpose of the control rods is to 
absorb some of the neutrons and thus stop them from causing further disintegrations. The control rods 
can be raised or lowered into the fuel rod bundle. When the control rods are lowered all the way into the 
fuel rod bundle, they absorb so many neutrons that the chain reaction essentially stops. When more heat 
is desired, the control rods are raised so they catch fewer neutrons, the chain reaction speeds up and more 
heat is generated. The control rods are operated in a fail-safe system so that power is necessary to hold 
them up; and during a power failure, gravity will pull the control rods down into shut off position. 




Hot coolant 



Control rods 
made of neutron 
absorbing 
material 



Uranium 
fuel rods 



Incoming coolant (cooler) 

U - 235 nuclei can capture neutrons and disintegrate more efficiently if the neutrons are moving slower 
than the speed at which they are released. Fission reactors use a moderator surrounding the fuel rods to 
slow down the neutrons. Water is not only a good coolant but also a good moderator so a common type of 
fission reactor has the fuel core submerged in a huge pool of water. This type of reaction is called a Light 
Water Reactor or LWR. All public electricity generating fission reactors in the United States are LWRs. 



587 



www.ckl2.org 



Reinforced concrete 
containment structure 



Steel contarment 
shell 



PRESSURIZED 

WATER REACTOR 

(PWR) 



Steam spins 
the turbine 




Note how everything 
is self contained. 
Nothing leaves the 
system except 
heat energy. 



Steam is cooled 
back to water 



You can follow the operation of an electricity-generating fission reactor in Figure ??. The reactor core is 
submerged in a pool of water. The heat from the fission reaction heats the water and the water is pumped 
into a heat exchanger container where the heated water boils the water in the heat exchanger. The steam 
from there is forced through a turbine which spins a generator and produces electricity. After the water 
passes through the turbine, it is condensed back to liquid water and pumped back to the heat exchanger. 

In the United States, heavy opposition to the use of nuclear energy was mounted in the late 1960s and 
early 1970s. Every environmentalist organization in the US opposed the use of nuclear energy and the 
constant pressure from environmentalist groups brought increased public fear and therefore, opposition. 
This is not true today; at least one environmental leader has published a paper in favor of nuclear powered 
electricity generation. 

In 1979, a reactor core meltdown at Pennsylvania's Three Mile Island nuclear power plant reminded the 
entire country of the dangers of nuclear radiation. The concrete containment structure (six feet thick walls 
of reinforced concrete), however, did what it was designed to do - prevent radiation from escaping into 
the environment. Although the reactor was shut down for years, there were no injuries or deaths among 
nuclear workers or nearby residents . Three Mile Island was the only serious accident in the entire history 
of civilian power plants (103 plants operating for 40 years) in the United States. There has never been a 
single injury or death due to radiation in any public nuclear power plant in the U.S. The accident at Three 
Mile Island did, however, frighten the public so that there has not been a nuclear plant built in the U.S. 
since the accident. 

The 103 nuclear power plants operating in the U.S. deliver approximately 19.4% of American electricity 
with zero greenhouse gas emission. There are 600 coal-burning electric plants in the US delivering 48.5% 
of American electricity and producing 2 billion tons of CO2 annually, accounting for 40% of U.S. CO2 
emissions and 10% of global emissions. These coal burning plants also produce 64% of the sulfur dioxide 
emissions, 26% of the nitrous oxide emissions, and 33% of mercury emissions. 4 



www.ckl2.org 



588 



Breeder Reactors 



U - 235 is the only naturally occurring fissile isotope and it constitutes less than 1% of naturally occurring 
uranium. It has been projected that the world's supply of U — 235 will be exhausted in less than 200 years. 
It is possible, however, to convert U — 238 to a fissionable isotope which will function as a fuel for nuclear 
reactors. The fissionable isotope is plutonium-239 and is produced by the following series of reactions. 



238 XT 1 
92 U + n 



239 

92 



, 239_ T 



239 

-I e+ 94 PU 



The final product from this series of reactions is plutonium-239 which has a half-life of 24, 000 years and 
is another nuclear reactor fuel. This series of reactions can be made to occur inside an operating nuclear 
reactor by replacing some of the control rods with rods of U - 238. As the nuclear decay process proceeds 
inside the reactor, it produces more fuel than it uses. It would take about 20 such breeder reactors to 
produce enough fuel to operate one addition reactor. The use of breeder reactors would extend the fuel 
supply a hundred fold. The problem with breeder reactors, however, is that plutonium is an extremely 
deadly poison and unlike ordinary fission reactors, it is possible for out-of-control breeder reactors to 
explode. None of the civilian nuclear power plants in the U. S. are breeder reactors. 

Radiation Detectors 

A variety of methods have been developed to detect nuclear radiation. One of the most commonly used 
instruments for detecting radiation and measuring the rate is the Geiger counter. 

A Geiger Counter. (Source: http://commons.wikipedia.Org/wiki/File:Geiger.png. GNU Free Doc- 
umentation) 



Cathode 



Ionizing Radiation 




The detecting component of the Geiger counter is the Geiger-Muller tube. This tube is a cylinder filled 
with an inert gas and it has a window in one end made of porous material that will not allow the inert gas 
to escape but will allow radiation particles to enter. A conducting wire extends into the center of the tube 



589 



www.cki2.0rg 



and is electrically insulated from the tube where it passes through the wall. The wire and walls are part of 
an electric circuit with a potential difference between the walls and the wire. Electric current does not flow 
because the circuit is not complete. The inert gas does not conduct electricity and therefore a gap exists 
in the circuit. When a radiation particle enters the tube through the window, the particle creates a line of 
ionized gas particles along its path through the tube. The line of ions does conduct electric current and 
an electric current will flow along the ionized path. The ions only exist for a very short period because the 
ions of inert gas will quickly regain the lost electrons and become atoms again - which do not conduct. The 
result is a very short burst of electric current whenever a radiation particle passes through the tube. The 
control box provides the electric potential for the tube and also provides some means for demonstrating 
the burst of current. Some machines simply make a clicking sound for each burst of current while others 
may provide a dial or a digital meter. 

Other methods used for the detection of nuclear radiation include, 1) scintillation counters - a screen 
coated with a material that gives off a small flash of light when struck by a particle, 2) cloud chambers 
(see image below) - a chamber of supersaturated gas that produce a condensation trail along the path of 
a radiation particle, and 3) bubble chambers - a chamber of superheated liquid that produces a trail of 
bubbles along the path of a radiation particle. 

Cloud Chamber showing vapor trails produced by sub-atomic particles. (Source: http://www.nasa.gov/ 
multimedia/imagegallery/image_f eature_928_prt .htm. Public Domain) 




Cloud chambers and bubble chambers have an additional value because the vapor trails or bubble trails left 
by the nuclear radiation particle are long-lasting enough to be photographed and therefore can be studied 
in great detail. 

Particle Accelerators 

In the early 1900's, the use of alpha particles for bombarding low atomic number elements became a 
common practice. Researchers found that the alpha particles were absorbed by the nuclei and a proton 
was ejected. This was the first artificially caused transmutation of one element into another. In order to 
continue these bombardments with alpha particles or protons, the speed of the bombarding particle had 
to be increased. Several machines were devised to accelerate the particles to the required speeds. 

The cyclotron was developed by Ernest Lawrence in 1930 and used to accelerate charged particles so they 
would have sufficient energy to enter the nuclei of target atoms. A cyclotron consists of two hollow half 

www.ckl2.org 590 



cylinders called "dees" because of their D-shapes. (Like a huge birthday cake which has been cut in half 
and the two halves separated by a short distance.) 

Cyclotron. 





The two "dees" have opposite charges with a potential difference of at least 50, 000 volts and the charges 
on the dees can be rapidly reversed so that each dee alternately becomes positive and then negative. 
The cyclotron also has a powerful magnetic field passing through it so that moving charged particles will 
be caused to travel in a curved path. Charged particles produced from a source in the center of the area 
between the dees are attracted first into one dee and then into the other as the charge on the dees alternate. 
As the particle moves back and forth in the dees, it is caused to follow a curved path due to the magnetic 
field. The motion of the particle is that of a spiral with ever increasing speed. As the circular path of the 
particle nears the outside edge of the cyclotron, it is allowed to exit through a window and strikes whatever 
target is placed outside the window. 

A linear accelerator is a long series of tubes which are connected to a source of high frequency alternating 
voltage. As the charged particles leave each tube, the charge on the tubes are altered so the particle is 
repelled from the tube it is leaving and attracted to the tube it is approaching. In this way, the particle 
is accelerated between every pair of tubes. The largest linear accelerators are located at the Fermilab in 
Illinois and at Stanford University. 

A number of the elements listed in the periodic table are not found in nature. These elements may never 
have been present on earth or since they have short half- lives, they may have originally been present but 
have completely decayed to more stable elements. These elements include all elements with atomic numbers 
greater than 92 plus technetium (#43) and promethium (#61). The transuranium elements (those with 
atomic numbers greater than 92) are all made man elements and many of them were produced in the 
cyclotron in the radiation laboratory at the University of California at Berkeley under the direction of 
Glenn Seaborg (Figure ??). Some very rare elements presence in the earth are assumed to be due not 
from the original material present in the earth but rather as daughter products of other disintegrating 
nuclei. 




591 



www.ckl2.org 



The following web site contains a history of the discovery (creation) of the transuranium 
elements and includes some of the reactions used to produce them. 

The Search for "Heavy" Elements (http://www.lbl.g0v/abc/wallchart/chapters/O8/O.html) 

Nuclear Medicine 

The held of nuclear medicine has expanded greatly in the last twenty years. A great deal of the expansion 
has come in the area of imaging. This section will focus on nuclear medicine involving the types of nuclear 
radiation introduced in this chapter. The x-ray imaging systems will not be covered. 

Radioiodine (1-131) Therapy involves imaging and treatment of the thyroid gland. The thyroid gland 
is a gland in the neck that produces two hormones that regulate metabolism. In some individuals, this 
gland becomes overactive and produces too much of these hormones. The treatment for this problem uses 
radioactive iodine (1-131) which is produced for this purpose in research hssion reactors or by neutron 
bombardment of other nuclei. 

The thyroid gland uses iodine in the process of its normal function. Any iodine in food that enters the 
bloodstream is usually removed by, and concentrated in the thyroid gland. When a patient suffering from 
an overactive thyroid swallows a small pill containing radioactive iodine, the 1-131 is absorbed into the 
bloodstream just like non-radioactive iodine and follows the same process to be concentrated in the thyroid. 
The concentrated emissions of nuclear radiation in the thyroid destroy some of the gland's cells and control 
the problem of the overactive thyroid. 

Smaller doses of 1-131 (too small to kill cells) are also used for purposes of imaging the thyroid. Once the 
iodine is concentrated in the thyroid, the patient lays down on a sheet of him and the radiation from the 
1-131 makes a picture of the thyroid on the him. The half- life of iodine- 131 is approximately 8 days so after 
a few weeks, virtually all of the radioactive iodine is out of the patient's system. During that time, they 
are advised that they will set off radiation detectors in airports and will need to get special permission to 
fly on commercial flights. 

Positron Emission tomography or PET scan is a type of nuclear medicine imaging. Depending on the 
area of the body being imaged, a radioactive isotope is either injected into a vein, swallowed by mouth, 
or inhaled as a gas. When the radioisotope is collected in the appropriate area of the body, the gamma 
ray emissions are detected by a PET scanner (often called a gamma camera) which works together with 
a computer to generate special pictures providing details on both the structure and function of various 
organs. PET scans are used to: 

• detect cancer 

• determine the amount of cancer spread 

• assess the effectiveness of treatment plans 

• determine blood flow to the heart muscle 

• determine the effects of a heart attack 

• evaluate brain abnormalities such as tumors and memory disorders 

• map brain and heart function 

External Beam Therapy (EBT) is a method of delivering a high energy beam of radiation to the precise 
location of a patient's tumor. These beams can destroy cancer cells and with careful planning, NOT kill 
surrounding cells. The concept is to have several beams of radiation, each of which is sub-lethal, enter 
the body from different directions. The only place in the body where the beam would be lethal is at the 
point where all the beams intersect. Before the EBT process, the patient is three-dimensionally mapped 
using CT scans and x-rays. The patient receives small tattoos to allow the therapist to line up the beams 

www.ckl2.org 592 



exactly. Alignment lasers are used to precisely locate the target. The radiation beam is usually generated 
with a linear accelerator. EBT is used to treat the following diseases as well as others: 

• breast cancer 

• colorectal cancer 

• head and neck cancer 

• lung cancer 

• prostate cancer 

Nuclear Weapons 

Nuclear weapons are of two basic types; fission bombs using supercritical masses of either U-235 or Pu-239 
and fusion bombs using heavy isotopes of hydrogen. The fission bombs were called atomic bombs (a 
misnomer since the energy comes from the nucleus) and fusion bombs are called thermonuclear bombs. 
Fission bombs use two or more subcritical masses of fissile material separated by enough distance that 
they don't become critical, and surrounded by conventional explosives. The conventional explosives are 
detonated to drive the subcritical masses of fissile material toward the center of the bomb and when these 
masses are slammed together, they form a supercritical mass and a nuclear explosion ensues. Hydrogen 
bombs (fusion) are detonated by using a small fission explosion to compress and heat a mass of deuterium 
or deuterium and tritium to the point that a fusion reaction ignites. 

Castle-Romeo nuclear explosion. (Source: http://en.wikipedia.Org/wiki/File:Castle_romeo2.jpg. 
Public Domain) 




Nuclear weapons are power-rated by comparison to the weight of conventional explosives (TNT) that would 
produce an equivalent explosion. For example, a nuclear device that produces an explosion equivalent to 
1,000 tons (2,000,000 pounds) of TNT would be called a 1-kiloton bomb. The atomic bomb detonated 
at Hiroshima near the end of WWII was a 13-kiloton bomb that used 130 pounds of U — 235. It has been 
estimated that this weapon was very inefficient and that less than 1.5% of the fissile material actually 
fissioned. The atomic bomb detonated at Nagasaki was a 21-kiloton weapon that used 14 pounds of 
Pu — 239. The two bombs set off in Japan would be considered very small bombs by later standards. 
Bombs that were tested later were measured not by kilotons but by megatons (million tons) of TNT. The 
largest bomb ever set off was a 50-megaton fusion weapon tested by the Soviet Union in the 1950s. 

The extensive death and destruction caused by these weapons comes from four sources. The tremendous 
heat released by the explosion heats the air so hot and so fast that the air expansion creates a wind in 
excess of 200 miles/hour - many times stronger than the strongest hurricane. The blast force from this 
wind completely destroys all but the strongest buildings for several miles from ground zero. The second 
source of damage is from fires ignited by the heat from the fireball at the center of the explosion. The 
fireball in the 50-megaton test was estimated to be four miles in diameter. The third source of injury 
is the intense nuclear radiation (primarily gamma rays) which are instantly lethal to exposed people for 

593 www.ckl2.org 



several miles. The final source of injury and possibly death conies from the radioactive fall-out which may 
be several tons of radioactive debris and may fall up to 300 miles or more away. This fall-out can cause 
sickness and death for many years. 

You should note that the distances given in the above material are guesses and estimates. The only actual 
data known are from the two (now considered small) bombs detonated in Japan and some above ground 
tests. The actual death and destruction from a 50-megaton bomb detonated in a heavily populated area 
is not known. 



Lesson Summary 



• The fission of U — 235 or Pu — 239 is used in nuclear reactors. 

• The critical mass is the amount of fissile material that will maintain a chain reaction. 

• Nuclear radiation also has many medical uses. 

Vocabulary 

control rods Control rods are made of chemical elements capable of absorbing many neutrons and are 
used to control the rate of a fission chain reaction in a nuclear reactor. 

cyclotron A cyclotron is a type of particle accelerator. 

fall out Fall out is radioactive dust hazard from a nuclear explosion, so named because it "falls out" of 
the atmosphere where it was spread by the explosion. 

fissile A fissile substance is a substance capable of sustaining a chain reaction of nuclear fission. 

fissionable A fissionable material is material capable of undergoing fission. 

Geiger counter A Geiger counter is an instrument used to detect radiation, usually alpha and beta 
radiation, but some models can also detect gamma radiation. 

isotope Nuclei with the same number of protons but different numbers of neutrons. 

linear accelerator A linear accelerator is a linear electrical device for the acceleration of subatomic 
particles. 

moderator A neutron moderator is a medium which reduces the velocity of fast neutrons; commonly 
used moderators are regular (light) water, solid graphite, and heavy water. 

nuclear pile A nuclear pile is a nuclear reactor. 

Further Reading / Supplementary Links 

• Nuclear Power VS. Other Sources of Power, Neil M. Cabreza, Department of Nuclear Engineering, 
University of California, Berkeley, NE-161 Report. Available at http://www.nuc.berkeley.edu/ 
thyd/nel61/ncabreza/sources . 

www.ckl2.org 594 



www.hps.org/publicinformation/ate/catlO.htnil 

www.sciencemag.org/cgi/content/full/309/5732/233 

www.hrd.qut.edu.au/toolkit/Faqs/radiation.jsp 

www.radiationnetwork.com/RadiationNetwork.htm 

http : //www . iaea . org/NewsCenter/Features/Chernobyl- 15 

http : //www . don . wa . gove/ehp/rp/f actsheets-pd/f slO 

http : //www . radscihealth . org/rsh/About 

http : //nrc . gov/reading-rm/doc-collections 

http : //www . world-nuclear . org/inf o/Chernobyl 

http : //www . iaea . org/NewsCenter/Features/Chernobyl- 15 

http : //www . nuclearweaponarchive . org/Russia/Tsarbomba . html 

http : //www . atomicarchive . com/effects/index . shtml 

Review Questions 

1. What is the primary physical difference between a nuclear electricity generating plant and a coal- 
burning electricity generating plant? 

2. What do the control rods in a nuclear reactor do and how do they do it? 

3. What is a breeder reactor? 

4. Name two types of particle accelerators. 

5. In the medical use of radioactivity, what does EBT stand for? 

6. Is it possible for a nuclear explosion to occur in a nuclear reactor? Why or why not? 



595 www.ckl2.org